Monografia - Processing Parameters of Rebar Steel A Case Study on Tempcore Graduation Project Report

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 61

Pr♦❝❡ss✐♥❣ P❛r❛♠❡t❡rs ♦❢

❘❡❜❛r ❙t❡❡❧
❆ ❈❛s❡ ❙t✉❞② ♦♥ ❚❡♠♣❝♦r❡
●r❛❞✉❛t✐♦♥ Pr♦❥❡❝t ❘❡♣♦rt
Pr❡♣❛r❡❞ ❜②✿
❊s❧❛♠ ▼♦❤❛♠❡❞ ❆❜❞✉❧♠❛✇❧❛
▼❛♥❛r ❆❤♠❡❞ ❊✐ss❛
▼❡♥♥❛t✉❧❧❛❤ ❆❤♠❡❞ ▼❛❤♠♦✉❞
▼♦❤❛♠❡❞ ❙♦❧✐♠❛♥ ❆❤♠❡❞
❙❛❧♠❛ ❙❤❛❛❜❛♥ ▼♦❤❛♠❡❞
❚❛r❡❦ ●❛❧❛❧ ▼♦❛✇❛❞

❙✉♣❡r✈✐s❡❞ ❜②✿
♣r♦❢✳ ❙❛♠❡❡r ■❜r❛❤✐♠ ❆❜❞✉❧❤❛❦❡❡♠
processing
parameters of
rebar steel A Case Study on Tempcore
by

Eslam Mohamed Abdulmawla


Manar Ahmed Eissa
Mennatullah Ahmed Mahmoud
Mohamed Soliman Ahmed
Nada Abdulhaleem Abdulhameed
Salma Shaaban Mohamed
Tarek Galal Moawad

to obtain the degree of Bachelor of Engineering at


Metallurgical and Materials Engineering department.
Faculty of Petroleum and Mining Engineering.
Suez University.

Project Supervisor: Prof. Samir Ibraheem


Project duration: February 8, 2020 – July 8, 2020
Examining Committee: Prof. Samir Ibraheem, Suez university, (supervisor)
Prof. Mohamed Amaar, Suez University
Dr. Adel Mohamed, Suez University
Dr. Marwa Abass, Suez University

An electronic version of this thesis is available at


❤tt♣s✿✴✴❞r✐✈❡✳❣♦♦❣❧❡✳❝♦♠✴❢✐❧❡✴❞✴✶❧❨t❨tq✾♣✷♥✸❈r❑✈r✲s❥rs✲q❡▲✇②❡t●❴❝✴✈✐❡✇❄✉s♣❂s❤❛r✐♥❣.
Preface
Reinforcement steel rebars, are used as a tension device in reinforced concrete to strengthen and hold the
concrete in tension, because Concrete is very strong in compression, but relatively weak in tension. There-
fore some requirements should be met in steel rebars such as: high yield strength, weldability, bendability
and ductility. Technological advances during the last few years in the field of deformed bars production
have helped in meeting all the above requirements together. Micro-alloying with Nb, V, Ti and B in com-
bination or individually, and thermo-mechanical treatment process are worth mentioning in this field. The
steel bars/wires for concrete reinforcement shall be manufactured by the process of hotrolling. It may be
followed by a suitable method of cooling and/or cold working. Technologies for producing such high quality
rebars involved either modifying the chemical composition of steel (alloying), or thermo-mechanical treat-
ment (Tempcore). Up today, the most diffused typology of steel for rebars in European constructions is Tem-
pCore. The TempCore process is characterized by two following phases of quenching and selftempering and
provides good strength and ductility towards moderate production costs, especially if compared to Micro-
Alloyed steels. Recent scientific works highlighted TempCore durability problems and drops of ductility and
dissipative capacity if exposed to aggressive conditions, both in the case of localized and uniform corrosion.
Two main routes can be pursued to solve or minimize corrosion effects: a ‘direct’ and an ‘indirect’ method.
The indirect method consists in the adoption of higher concrete classes, thicker concrete covers, higher di-
ameter rebars, etc. By this way, corrosion effects are minimized, but the source still exists: this is the method
proposed by European standards. The other possibility is to avoid corrosion initiation and propagation by se-
lecting opportune materials less exposed to durability problems. During the last years the scientific interest
in the possibility to enhanced steels for rebars strongly increased, investigation on the possibility to adopt, for
civil constructions, Dual-Phase (DP) steels is performed. DP steels are widely used in the automotive sector
since characterized by excellent ductile properties and improved durability performance due to their specific
micro-structure, characterized by a ferrite matrix in which martensite is directly embedded.
Preliminary investigations concerning RC-DP structures highlighted the improved performance of RC re-
spect to RC-TEMP ones. The higher ductility of DP steels can lead to greater ductility of RC elements: DP
steel enables attaining better performance in the post-elastic field thanks to the enlargement of the plastic
zone within the element. Results of mechanical investigations on DP rebars highlighted average values of the
hardening ratio (σU T S /σY ) around 2.0 and high elongation. The use of DP steel in constructions is, currently,
limited by industrial aspects: the achievement of DP bars need the modification of steel reinforcing plants
and of the production process. Costs shall be, otherwise, limited to allow the commercialization of the prod-
uct compared to ordinary reinforcing steels. The present project shows the procedure to produce DP rebars
using old TEMPCORE rebars, including selection of chemical composition, analysis of the thermal process,
and investigation of the production aspects and microstructural and mechanical investigations

iii
Acknowledgment
First of all, we thank Allah the most beneficent, the most merciful for endowing us with health and knowledge
to work on this project. We wish to express our deepest gratitude to Prof. Samir Abd-Elhakim our supervisor
and mentor for his encouragement, guidance, effort and support from the initial to the final steps enabling us
to understand and achieve the project main aims. Great thanks to prof. Rashad Ramdan for providing us with
the Insights needed in our project and for helping us tackle many difficulties that faced us during the project.
Moreover, we would like to thank all of the faculty’s board and all teaching assistants and EZDK Co. for their
Assisting us with carrying out the necessary experiments, We also pay our special regards to everyone who
supported and taught us through our whole educational life. . .

Graduation Project Report


Suez University, August 2020

v
Contents

✶ ■♥tr♦❞✉❝t✐♦♥ ✶

1.1 Iron Carbon System . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1


1.2 Phase transformations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
1.2.1 Austenite-Ferrite Transformation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
1.2.2 Martensite Transformation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
1.2.3 Effect of alloying elements on martensite formation . . . . . . . . . . . . . . . . . . . . 6
1.3 Introduction to Thermal Heat Treatment Processes . . . . . . . . . . . . . . . . . . . . . . . . 6
1.3.1 Annealing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
1.3.2 Process Annealing and Stress Relief . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
1.3.3 normalizing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
1.3.4 Spheroidizing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
1.3.5 Hardening . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
1.3.6 Continuous Cooling Transformation Diagrams . . . . . . . . . . . . . . . . . . . . . . 10
1.3.7 Austenitizing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
1.3.8 Quenching . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
1.3.9 Tempering . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14

✷ ❚❊▼P❈❖❘❊
➞ ✱ ❆ ▲♦✇ ❈♦st ❍✐❣❤ ❙tr❡♥❣t❤ ❘❡❜❛r✳ ✷✶

2.1 A Breif Historical Background . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21


2.2 The Tempcore Process . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
2.2.1 Theory of the TEMPCORE Process . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
2.2.2 Fundamental Mechanisms of Tempcore Hardening . . . . . . . . . . . . . . . . . . . . 26
2.2.3 Surface Heat Exchanges . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
2.2.4 Heat Conduction in the Bar . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
2.2.5 Volume Percentage of Martensite (p m ): . . . . . . . . . . . . . . . . . . . . . . . . . . 28
2.2.6 Tempering Temperature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
2.2.7 Control Function of the Process . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
2.2.8 Mechanical Properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31
2.2.9 Central Region of the Bar. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
2.2.10 Overall Yield Strength of the Tempcore Bar. . . . . . . . . . . . . . . . . . . . . . . . . 34
2.2.11 Practical Conclusions of the Theoretical Study . . . . . . . . . . . . . . . . . . . . . . . 35
2.3 Characteristics of Tempcore Reinforcing Steels . . . . . . . . . . . . . . . . . . . . . . . . . . 35
2.3.1 Type of steel . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35
2.3.2 Metallurgical phases and microstructures . . . . . . . . . . . . . . . . . . . . . . . . . 35
2.3.3 Effects of process parameters and steel composition. . . . . . . . . . . . . . . . . . . . 35
2.3.4 Tensile properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36
2.3.5 Weldability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37
2.3.6 Economic aspects . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38
✸ ❉✉❛❧ P❤❛s❡ ❙t❡❡❧ ✸✾

3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
3.1.1 Definition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
3.1.2 Microstructure of DP Steels . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40
3.1.3 DP Steel Production . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40
3.2 Microstructure Development during CA . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41
3.2.1 Inheritance from the hot rolling process . . . . . . . . . . . . . . . . . . . . . . . . . . 41
3.2.2 Cold Rolling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43
3.2.3 Heating & Soaking . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43
3.2.4 Cooling & Quenching . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43
3.2.5 Tempering . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43

vii
viii Contents

3.3 Mechanical Properties of DP Steels . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43


✹ Pr❛❝t✐❝❛❧ ❲♦r❦ ✹✺
4.1 Practical Work Objectives . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45
4.2 Work Procedure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45
4.2.1 Test material . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46
❇✐❜❧✐♦❣r❛♣❤② ✺✶
1
Introduction
Steel remains the most successful and cost-effective of all materials, with more than a billion tonnes being
consumed annually in improving the quality of life. This book attempts to explain why steels continue to take
this pre-eminent position, and examines in detail the phenomena whose exploitation enables the desired
properties to be achieved. One reason for the overwhelming dominance of steels is the endless variety of
microstructures and properties that can be generated by solid-state transformation and processing.

1.1. Iron Carbon System


At least three allotropes of iron occur naturally in bulk form, body-centred cubic (bcc, α, ferrite), face-centred
cubic (fcc, γ , austenite) and hexagonal close-packed (hcp, ǫ). The phase β in the alphabetical sequence
α,β,γ,δ... is missing because the magnetic transition in ferrite was at one time incorrectly thought to be the β
allotrope of iron, responsible for hardening when the iron is quenched. Nevertheless, it is true that α-iron is
not strictly cubic in its ferromagnetic state below the Curie temperature. This is because the magnetic spins
are aligned, say along the z axis, so that rotations about the x or y axes must be combined with time reversal
to preserve the directions of the spins. The magnetic point group becomes tetragonal. It follows that the
ferrite structure below the Curie temperature is tetragonal, but the extent of tetragonality is very small and
neglected in most experiments on steels containing complex microstructures where X-ray diffraction peaks
are too broad to detect small differences in lattice parameters.
A study of the constitution and structure of all steels and irons must first start with the iron-carbon equi-
librium diagram. Many of the basic features of this system (Fig. 1.1) influence the behavior of even the most
complex of alloy steels. For example, the phases found in the simple binary Fe-C system persist in complex
steels, but it is necessary to examine the effects alloying elements have on the formation and properties of
these phases. The iron-carbon diagram provides a valuable foundation on which to build knowledge of both
plain carbon and alloy steels in their immense variety. It should first be pointed out that the normal equilib-
rium diagram really represents the metastable equilibrium between iron and iron carbide (Fe3 C, cementite).
Cementite is metastable relative to the equilibrium between iron and graphite. Although graphite occurs ex-
tensively in cast irons(2–4 wt% C), it usually is difficult though not impossible to achieve this equilibrium in
steels (0.03–1.5 wt% C) without annealing for a long time. Therefore, the metastable equilibrium between
iron and iron carbide should be considered, because it is relevant to the behaviour of most steels in practice.
The larger phase field of γ-iron (austenite) compared with that of α-iron (ferrite) reflects the much greater
solubility of carbon in γ-iron, with a maximum value of just over 2 wt% at 1147◦ C when it is in equilibrium
with liquid and cementite (E, Fig. 1.1). This relatively high solubility of carbon in γ-iron is of extreme im-
portance in heat treatment, when solution treatment in the γ-region followed by rapid quenching to room
temperature allows a supersaturated solid solution of carbon in iron to be formed. The α-iron phase field is
severely restricted, with a maximum carbon solubility of 0.02 wt% at 727◦ C when the ferrite is in equilibrium
with austenite or cementite (‘P’, Fig. 1.1), so over the carbon range encountered in steels from 0.05 to 1.5 wt%,
α-iron is normally associated with iron carbide in one form or another. Similarly, the δ-phase field is very
restricted between 1390 and 1538◦ C and disappears completely when the carbon content reaches 0.5 wt%.
There are several temperatures or critical points in Fig. 1.1 which are important, both from the basic and
from the practical point of view. Firstly, there is the Ae 1 temperature at which the eutectoid reaction occurs

1
2 1. Introduction

Figure 1.1: The iron-carbon diagram showing the equilibrium between the liquid, ferrite, austenite and cementite phases; the dotted
line marks the Curie temperature of ferrite. The eutectoid composition is 0.76C wt%. Cementite has the composition ≈6.67 wt% carbon
corresponding to 25 at%. However, the composition of cementite becomes iron-rich at high temperatures. As a consequence, ferrite
precipitates within the cementite produced by the high-temperature eutectic reaction, on cooling to lower temperatures.

(PS-K), which is 727◦ C in the binary diagram. Secondly, there is the Ae 3 temperature when α-iron transforms
to γ-iron. For pure iron this occurs at 912◦ C, but the transformation temperature is progressively lowered
along the line (GS) by the addition of carbon. The third point is Ae 4 at which γ-iron transforms to δ-iron,
1390◦ C in pure iron, but this is raised as carbon is added. The Ae 2 point is the Curie temperature when
ferritic iron changes from the ferro to the para-magnetic condition or vice-versa. This temperature is 769◦ C
for pure iron, but there is only a very slight change in crystal structure involved because ferromagnetic ferrite
is body-centred tetragonal with the difference between the ’c’ and ’a’ lattice parameters being small enough
to be neglected in most circumstances. The Ae 1 , Ae 3 and Ae 4 points are easily detected by thermal analysis
or dilatometry during cooling or heating experiments conducted at rates slow enough to achieve equilibrium.
When equilibrium experiments are not conducted, two further values are associated with each equilibrium
temperature, i.e., A c for heating (chauffage) and A r for cooling (refroidissement). It is emphasized that the
A c and A r values will be sensitive to the rates of heating and cooling, as well as to the presence of alloying
elements.

1.2. Phase transformations


1.2.1. Austenite-Ferrite Transformation
Under equilibrium conditions, pro-eutectoid ferrite will form in iron-carbon alloys containing up to 0.76 wt%
carbon. The reaction occurs at 912◦ C in pure iron, but takes place between 912◦ C and 727◦ C in iron-carbon
alloys. However, by supercooling the austenitic state to temperatures below the eutectoid temperature, ferrite
can be formed in reasonable time periods down to temperatures as low as 600◦ C. There are pronounced
morphological changes as the transformation temperature is lowered, of a type that applies to both hypo-
and hyper-eutectoid phases, although in each case there will be variations due to the precise crystallography
of the phases involved. For example, the same principles apply to the formation of cementite from austenite,
but it is not difficult to distinguish ferrite from cementite morphologically.
The transformation of austenite in steels can be studied during continuous cooling using various phys-
ical measurements, e.g. dilatometry, thermal analysis, electrical resistivity, hot-stage microscopy etc., how-
ever, the results obtained are very sensitive to the cooling rate used. Davenport and Bain first introduced
the isothermal transformation approach in combination with good microscopy to show that by studying the
reaction isothermally at a series of temperatures, a characteristic time-temperature-transformation or TTT
1.2. Phase transformations 3

curve can be obtained for each particular steel. In their simplest form, these transformation curves have a
well-defined ‘C’ shape (Fig. 1.2), where there is a temperature at which the reaction proceeds most rapidly,
slowing down both at higher and at lower temperatures. That temperature, at which the reaction rate is
fastest, is often referred to as the nose of the C-curve. This can be explained in general terms as follows. For
a eutectoid steel transformed close to the eutectoid temperature, the degree of undercooling, ´T, is low so
the driving force for the transformation is small. However, as ´T increases, so does the driving force mak-
ing the reaction faster, until the maximum rate at the nose of the curve. Below this temperature, the driving
force for the reaction continues to increase, but the reaction is now impeded by the slow diffusivity of the
rate-controlling element, which in plain carbon steels may be carbon or iron, depending on the transfor-
mation conditions. One of the simplest examples of a TTT curve is that for a steel that has a composition
approximating the eutectoid carbon content. In Fig. 1.2 the beginning and end of transformation over a wide
temperature range is plotted to produce two curves making up the diagram. When the carbon content of
the steel is lowered, the ferrite reaction will also take place and this is represented by another curve which is
frequently imposed on the same diagram, and which normally precedes the pearlite reaction. Similarly, the
cementite reaction can be recorded in hyper-eutectoid steels. The TTT curve strictly applies to the nucleation
and growth of one phase in austenite, but at the lower temperatures other constituents can appear, e.g. the
displacive transformation products Widmanstätten ferrite, bainite, martensite. These have quite different
characteristics to ferrite and pearlite that form by a reconstructive mechanism that entails the diffusion of all
atoms.

Figure 1.2: TTT diagram for a Fe−0.89 C−0.29 Mn wt% steel. The curves here represent overlapping information from the separate
C-curves of reconstructive and displacive transformations because of the rapid transformation rate and the technique used (adapted
from US Steel Co., Atlas of Isothermal Diagrams). The carbon concentration exceeds 0.76 wt% so some proeutectoid cementite is
possible but not illustrated here.

1.2.2. Martensite Transformation


The quenching to room temperature of austenite in a steel can lead to the formation of martensite, a very
hard phase in which the carbon, formerly in solid solution in the austenite, remains in solution in the new
phase. The velocity of the interface is greater than the ability of carbon to diffuse away into the austenite:
γ
DC
v α′ > (1.1)
λ
4 1. Introduction

where λ is the distance between the interstices in which carbon resides, so the term on the right is the speed
with which the carbon can diffuse. The carbon is said to be trapped within the martensite. Unlike ferrite
or pearlite, martensite forms by a deformation of the austenite lattice without any diffusion of atoms. The
deformation causes a change in the shape of the transformed region, consisting of a large shear and a volume
expansion. Martensite is, therefore, often referred to as a diffusionless, shear transformation, which is highly
crystallographic in character because it is generated by a specific deformation of the austenite. When the
formation of martensite is constrained by its surroundings, it forms as thin plates or laths in order to minimize
the strain energy due to the deformation.
The martensite reaction in steels normally is said to occur athermally, i.e. the fraction transformed de-
pends on the undercooling below a ‘martensite-start temperature, MS . The extent of transformation does not
seem to depend on time, as expressed in the Koistinen and Marburger Equation 1.2 equation which describes
the progress of transformation below M S :

1 − VVα′ = exp{β(M S − TQ )} where β ≈ −0.011. (1.2)

VVα is the volume fraction of martensite and TQ the temperature below M S to which the sample is cooled. This
apparent athermal character is a pragmatic description – the absence of time in Equation 1.2 is a consequence
of very rapid transformation on time scales much shorter than observable by ordinary perception. In such
cases, the reason why the undercooling below M S must be increased in order to achieve further transforma-
tion is because the easy nuclei are triggered at small undercoolings, followed progressively by the remaining
less potent nuclei as M S − TQ increases. Martensite is, like all other phases that evolve from austenite, a first
order transformation involving nucleation and growth. Both of these processes require thermal activation,
but the activation energy is small because there is no diffusion required during martensitic transformation,
and because the interfacial structure is conducive to easy glide.

Figure 1.3: time-temperature-transformation diagram for the isothermal formation of martensite in Fe-0.05C-23.6Ni-3.3Mn-0.2Cu
wt%. These are the first experiments to reveal the role of thermal activation in martensitic transformation. Selected data from
Kurdjumov and Maximova

Nevertheless, the fact that thermal activation is required to overcome the small barriers means that there
is a time dependence of transformation, which can be detected readily if the transformation occurs at cryo-
genic temperatures. Using a richly alloyed steel, Kurdjumov and Maximova first demonstrated the isother-
mal formation of martensite, following a classical C-curve behaviour, Fig. 1.3. On the same reasoning, the
transformation to martensite can also be suppressed completely by cooling sufficiently rapidly to very low
temperatures; a Fe-33Ni wt% alloy cooled to 4 K remained fully austenitic but then began to transform into
martensite on warming to about 42 K . Nevertheless, for most steels where the M S temperature is well above
ambient, the kinetics of transformation appear independent of time and follow (Equation 1.2), from which
it is evident that some austenite remains untransformed when TQ is set to room temperature. This is re-
ferred to as retained austenite. It is also clear that there is no martensite-finish temperature, M F , but for
convenience the latter is frequently defined at the point where 95% of the martensitic transformation is com-
pleted. Martensite is not restricted to steels although its technological importance in steels is unsurpassed.
Table 1.1 lists a variety of materials which exhibit martensitic transformation, together with M S temperatures
1.2. Phase transformations 5

Table 1.1: The temperature M S at which martensite first forms on


cooling, and the approximate Vickers hardness of the resulting
martensite for a number of materials

Composition M S /K Hardness/HV
ZrO2 1200 1000
Fe – 31 Ni – 0.23 C wt% 83 300
Fe – 34 Ni – 0.22 C wt% <4 250
Fe – 3 Mn – 2 Si – 0.4 C wt% 493 600
Cu – 15 Al at% 253 200
Ar – 40 N2 at% 30
In – 25 Tl at% 125
In – 29 Tl at% 193
Gd – 35 Ce at% 763

and hardness values. To obtain martensite, it usually is necessary for the steel to be cooled from the austenite
phase field at a rate which is sufficiently fast to avoid all other solid-state transformations such as ferrite and
pearlite. This cooling rate can be very high for plain carbon steels but quite slow for a heavily alloyed steel
containing large concentrations of austenite stabilizing solutes. Martensite can form at very low tempera-
tures, where diffusion, even of interstitial atoms, is not conceivable over the time period of the experiment.

Figure 1.4: Definition of the habit plane, the broad interface between the martensite and austenite. The interface is flat during
unconstrained transformation but curved when the displacements due to martensitic transformation are constrained by the
surrounding material. The average plane of the martensite is then the habit plane, which is parallel to the flat plane of the
unconstrained transformation

Table 1.1 gives typical values of the martensite-start temperature for a variety of materials. Martensite plates
can grow at speeds which are a substantial fraction of the speed of sound in the steel, some 1100 ms−1 Such a
high rate of growth is inconsistent with diffusion during transformation.Transformations which involve dif-
fusion are much slower – the fastest recorded solidification rate is about 80 ms−1 in pure nickel, illustrating
the difference between coordinated and uncoordinated atomic motions. The difference is even greater when
the random diffusional jumps happen in the solid state; thus the speed of a massive transformation has been
reported to be only 0.03 ms−1 .
The chemical composition of martensite can be measured and shown to be identical to that of the parent
austenite. These observations also demonstrate convincingly that martensitic transformations are diffusion-
less.
6 1. Introduction

1.2.3. Effect of alloying elements on martensite formation

Figure 1.5: The effect of carbon on the martensite-start (M S ) and martensite-finish (M F ) temperatures. The latter is not well defined
but can be set to correspond to an arbitrary 95% of transformation. Data from Petty.

Most alloying elements which enter into solid solution in austenite lower the M S temperature, with the
exception of cobalt and aluminium. However, the interstitial solutes carbon and nitrogen have a much larger
effect than the metallic solutes. The effect of carbon on both M S and M F is shown in Fig. 1.5, from which it can
be seen that 1wt% of carbon lowers the M F by over 300◦ C. Note that above 0.7 wt% C the M F temperature is
below room temperature and consequently higher carbon steels quenched into water will normally contain
substantial amounts of retained austenite. The relative effect of other alloying elements is indicated in the
following empirical relationship due to Andrews:

M S (◦C ) = 539 − 423wC − 30.4w Mn − 17.7w N i − 12.1wC r − 7.5w Mo (1.3)

where w i represents the weight percent of the element identified by the subscript and the limits of the
data on which this equation is based are:

C Mn Si Ni Cr Mo
Minimum / wt% 0.11 0.20 0.11 0.00 0.00 0.00
Maximum / wt% 0.55 1.67 1.74 5.04 3.34 1.00

The uncertainty with relationships such as Equation (Equation 1.3) is about ±20◦ C, which is associated largely
with the noise in conducting experiments that measure the M S temperature. The assumption that M S is a
linear function of solute concentrations may not be justifiable, and more complex expressions that involve
products of concentrations are certainly arbitrary; neural network methods are much better in this respect
because the non-linear functions appropriate to represent variations in data are derived rather than assumed.
It is not surprising therefore, that the gradient of the curve in Fig. 1.5 is different from that implied by Equation
(Equation 1.3), since the relationship between M S and carbon is not independent of the other solutes.

1.3. Introduction to Thermal Heat Treatment Processes


One of the primary advantages of steels is their ability to attain high strengths through heat treatment while
still retaining some degree of ductility. This ability of steels to be strengthened is a direct consequence of the
amount of carbon present (Fig. 1.6). As the carbon content is increased, higher strength levels are obtainable.
Although the ductility decreases with increasing strength, it is still high enough to satisfy most engineering
applications. The type of heat treatment used also influences the properties. As shown in Fig. 1.7, a steel
hardened by heating into the austenite field, followed by quenching and then tempering, has a much higher
strength than one subjected to either a normalizing or spheroidizing heat treatment. Heat treatments can be
used to not only harden steels but also to provide other useful combinations of properties, such as ductility,
formability, and machinability. The various heat treatment processes covered in this section include anneal-
ing, stress relieving, normalizing, spheroidizing, and hardening by quenching and tempering. In all of these
processes, the steel is heated fairly slowly to some predetermined temperature and then cooled.
1.3. Introduction to Thermal Heat Treatment Processes 7

Figure 1.6: Effect of carbon content on steel strength. UTS, ultimate tensile strength; YS, yield strength

It is the temperature and the cooling rate that determine the resultant structure of the steel and hence
the mechanical properties, as illustrated in the time temperature transformation (TTT) diagram in Fig. 1.8.
The final structure is independent of the heating rate, provided it has been slow enough for the steel to reach
structural equilibrium at the maximum temperature. However, the subsequent cooling rate, which deter-
mines the nature of the final structure, is critical and may vary between slow furnace cooling to rapid cooling
by quenching in water.

Figure 1.7: Effect of heat treatment on hardness of steel


8 1. Introduction

1.3.1. Annealing
The term annealing is a heat treatment in which a metal or alloy is heated to a certain high temperature and
held for a certain period of time, followed by slow cooling. In steels, annealing usually has the meaning of
a heat treatment with furnace cooling from the austenitizing range shown in Fig. 1.9. Annealing is used to
reduce hardness, obtain a relatively near stable microstructure, refine grain size, improve machinability, and
facilitate cold working. For hypoeutectoid steels, full annealing consists of heating to 10 to 40 ◦ C above the
A 3 temperature, and for hypereutectoid steels, heating above the A 1 temperature and followed, in both cases,
by very slow cooling. As a result of the very slow cooling rates, the microstructure consists of coarse ferrite or
coarse ferrite plus pearlite, depending on the carbon and alloy content of the steel. Isothermal annealing is
often more efficient for small parts because, instead of furnace cooling, the parts can be transferred between
constant-temperature molten salt baths. The steel is also protected from scaling and oxidation by the molten
bath. These salts are various mixtures of compounds such as alkali metal hydroxides and nitrates that are
chosen for particular temperature ranges. Baths of molten lead were commonly used before concerns about
the dangers of lead to human health and the environment became widespread

Figure 1.8: Effect of cooling rate on microstructure.

1.3.2. Process Annealing and Stress Relief


As the hardness of steel increases during cold working, ductility decreases, and additional cold reduction
becomes so difficult that the material must be annealed to restore its ductility. Such annealing between pro-
cessing steps is referred to as in-process or simply process annealing. It may consist of any appropriate treat-
ment; however, in most instances, a subcritical treatment is adequate and the least costly. The term process
annealing without further qualification usually refers to an in-process subcritical anneal. Process annealing
usually consists of heating steel to a temperature just below the A 1 eutectoid temperature for a short time.
This provides stress relief, makes the steel easier to form, and is applied to low-carbon cold-rolled sheet steels
to restore ductility. The temperatures used range from 550 to 650◦ C. slow cooling is not essential for process
annealing, since any cooling rate from temperature below A 1 will not affect the microstructure or hardness.
1.3. Introduction to Thermal Heat Treatment Processes 9

Although recrystallization can occur due to the stored energy from cold working, there are no phase changes,
and the ferrite and cementite constituents remain the same throughout the process. Process annealing is
generally carried out in either batch-type or continuous furnaces

1.3.3. normalizing
Steel is normalized by heating 30 to 45◦ C into the austenite phase field above the A 3 temperature, somewhat
higher than those used by annealing, followed by cooling at a medium rate. For carbon steels and low-alloy
steels, normalizing means air cooling. Many steels are normalized to establish a uniform ferrite plus pearlite
microstructure along with a uniform grain size. The faster cooling rate during normalizing results in a much
finer pearlitic structure, which is harder and stronger than the coarse pearlite produced by full annealing.
Steel is normalized to refine grain size, make its structure more uniform, make it more responsive to harden-
ing, and to improve machinability. When steel is heated to a high temperature, the carbon can readily diffuse,
resulting in a reasonably uniform composition from one area to the next. The steel is then more homoge-
neous and will respond to the heat treatment more uniformly. The properties of normalized steels depend on
their chemical composition and on the cooling rate, with the cooling rate being a function of the size of the
part. Although there can be a considerable variation in the hardness and strengths of normalized steels, the
structure usually contains fine pearlite.

Figure 1.9: Steel heat treating ranges.

1.3.4. Spheroidizing
To produce a steel in its softest possible condition with minimum hardness and maximum ductility, it can
be spheroidized by heating just above or just below the A 1 eutectoid temperature and then holding at that
temperature for an extended period of time. Spheroidizing can also be conducted by cyclic processing, in
which the temperature of the steel is cycled above and below the A 1 line. This process breaks down lamellar
pearlite into small pieces and forms small cementite spheroids through diffusion in a continuous matrix of
ferrite (Fig. 1.10). Surface tension causes the carbide particles to develop a spherical shape. Since a fine
initial carbide size accelerates spheroidization, the steel is often normalized prior to spheroidizing. Another
10 1. Introduction

option is to start with a martensitic structure that produces a very uniform dispersion of cementite spheroids,
because carbon is more uniformly distributed in martensite than in lamellar pearlite. The cementite lamellae
must first dissolve and then redistribute the carbon as spheroids, whereas cementite spheroids can form
directly from martensite.

1.3.5. Hardening
Steels are hardened by austenitizing, quenching, and then tempering to their final hardness. Since the com-
position of steels varies quite a bit, it is important to:
a) understand the maximum section thickness that can be hardened in a specific quench medium (e.g.,
water or oil)
b) realize the large variations in final strength and ductility that can be obtained by tempering at different
temperatures.
The generalized procedure for quenching and tempering is illustrated in Fig. 1.11. It should be noted that
tempering is not used to harden steel. Steel is hardened by austenitizing and quenching. Tempering is con-
ducted to restore a portion of the ductility that was lost during hardening, and often, appreciable softening is
produced by tempering. The carbon content is critical to the ability to harden steel. Since ductility decreases
with increasing carbon content, the carbon content is held to approximately 0.45 wt% in many engineering
steels. However, when wear resistance is required, for example, in tool and die steels, it may be increased to
over 1.0 wt%. The addition of alloying elements shifts the nose of the TTT diagram to the right, thus allowing
thicker sections to be hardened or allowing less drastic quenches. The effect of alloying elements and sec-
tion size on hardenability is illustrated in Fig. 1.12. In this example, both the 1040 and 4140 steels contain
nominal carbon contents of 0.40%, and yet, due to the alloying elements in 4140, 4140 hardens to a much
greater depth. However, as the diameter of a bar of 4140 is increased from 5 to 10 cm, the depth of hardening
(hardenability) decreases. Some more highly alloyed steels have TTT diagrams shifted so far to the right that
they will form fully martensitic structures in fairly thick sections by air cooling to room temperature. There-
fore, while some alloying elements may not directly increase the hardness of martensite significantly, they do
increase the hardenability, which is the depth from the surface of the martensite microstructure that can be
produced in steel of a given carbon content during quenching.

Figure 1.10: Microstructure of spheroidized steel.

1.3.6. Continuous Cooling Transformation Diagrams


Isothermal transformation diagrams (TTT curves) provide valuable information about the kinetics of the
transformation of austenite and are useful tools for developing heat treatments. However, in practical heat
1.3. Introduction to Thermal Heat Treatment Processes 11

treatments, where the size of the part being treated is usually significant, the temperature and the rate of
change of temperature vary with position within the part being heat treated. Also, there are heat treatments
in which the steel is continuously cooled. Therefore, for industrial applications, continuous cooling transfor-
mation (CCT) diagrams are more useful than TTT diagrams. During continuous cooling, part of the time is
spent at higher temperatures where the transformation rates are slower. Therefore, the time for transforma-
tion is longer for continuous cooling than for isothermal cooling. The lines representing the start of pearlite
and bainite formation are shifted to longer times and lower temperatures. Since alloying elements tend to
delay the formation of pearlite more than they delay bainite formation, bainite formation can occur during
continuous cooling. The TTT and CCT diagrams for 4340 are compared in Fig. 1.13 and Fig. 1.14, respectively.
There are two main types of CCT diagrams. The first type, shown in Fig. 1.14, has axes that show the tem-

Figure 1.11: Hardening heat treatment for steels

perature of the steel on the ordinate axis and time in seconds plotted on a logarithmic scale on the abscissa.
Lines are drawn on the diagram to show the cooling path in terms of the variation of temperature with time.
By considering such a cooling path and traveling along it downward from the austenite, the start of a trans-
formation is encountered. The location of this point is marked on the diagram and joined to similar points
on other cooling paths to create curved lines that denote the initiation of the decomposition of the austen-
ite. Similar lines that show the progress of the transformation in terms of the percent completed are often
placed between the start and finish curves. Diagrams of this type are determined experimentally by dilatom-
etry. Most commercial martensitic steels contain alloying additions intended to suppress the formation of
other constituents, namely ferrite, pearlite, and bainite, during continuous cooling. This means that these
constituents form at slower cooling rates, allowing martensite to form at the faster cooling rates, for example,
during oil and water quenching. Most of the conventional alloying elements in steel promote hardenability.
For example, the 4340 steel shown has significant levels of carbon, manganese, nickel, copper, and molybde-
num to promote hardenability. The second type of CCT diagram retains temperature as the ordinate axis but
defines the abscissa in terms of diameters of steel bars cooled in air, oil, and water, respectively, as shown in
Fig. 1.15. These quenching media are commonly used in industrial practice and extract heat from the surface
of a bar at different rates,with the water quench being the most rapid. Since different oil quenchants do not
extract heat at the same rates, the diagrams provide only qualitative guidelines. There are also variations in
chemical composition among steels of the same designation, which can affect the transformations.

1.3.7. Austenitizing
During austenitization, the steel is heated into the austenite (γ) field and held for a sufficient period of time
to dissolve most of the carbides and put them into solid solution. As previously shown in Fig. 1.9, the temper-
ature required for austenitization is a function of the carbon content in carbon steel. With increasing carbon
contents, the temperature decreases along the A 3 line to a minimum value at A 1 , the eutectoid composition
12 1. Introduction

Figure 1.12: Effect of alloying and section size on hardenability of steel alloys 1040 and 4140

(0.8% C), and then increases again along the A cm line. The first stage in the formation of austenite is the
nucleation and growth of austenite from pearlite (ferrite+Fe3 C). Even after the complete disappearance of
pearlite, some carbides will remain in the austenite.
The formation of homogeneous austenite is accelerated by increasing the temperature and increasing the
fineness of the initial carbide particles. To minimize the time for austenitization, the temperature used is ap-
proximately 55◦ C above the minimum temperature for 100% austenite, and the time is approximately 1 h per
2.5 cm of thickness. However, it is also important to keep the austenitization temperature as low as possible
to reduce the tendency toward cracking and distortion, minimize oxidation and decarburization, and min-
imize grain growth. The temperatures needed for obtaining 100% austenite in hypereutectoid steels can be
quite high; however, austenite suitable for hardening in these steels can be obtained at approximately 770◦ C.
The small amount of undissolved carbides dispersed in the austenite has little effect on the final mechanical
properties. High temperatures are also required for very low carbon steels, but carbon steels containing less
than 0.2% C respond poorly to quenching and are seldom used in the quenched and tempered condition.

1.3.8. Quenching
While the face-centered cubic (fcc) austenite that forms during austenitization is capable of dissolving as
much as 2 wt% C, only a small fraction of carbon can be retained in the lower temperature body-centered
cubic (bcc) ferrite. If the steel is slowly cooled from the austenitization temperature, carbon atoms are re-
jected as the fcc austenite transforms to the bcc ferrite, and alternating layers of ferrite and cementite form
pearlite through a nucleation and growth process. However, if the steel is rapidly cooled from the austeniti-
zation temperature (quenched), the carbon does not have time to diffuse out of the austenite microstructure
when it transforms into the body-centered tetragonal (bct) structure called martensite. Thus, the objective
of the quenching process is to cool at a sufficient rate to form martensite. The distortion of the bct structure
results in high strength and hardness of the quenched steel. As previously shown in Fig. 1.11, the steel must
be cooled past the nose of the isothermal transformation diagram to form 100% martensite. Martensite does
not form until it reaches the martensite start temperature (M S ) and is complete after it is cooled below the
1.3. Introduction to Thermal Heat Treatment Processes 13

Figure 1.13: Isothermal transformation diagram for 4340 steel. Austenitized at ASTM grain size 6–7.

martensite finish temperature (M F ). The addition of alloying elements increases the hardenability of steels
by moving the nose of the isothermal transformation diagram to the right, allowing slower cooling rates for
alloy steels to form martensite. However, alloying elements also depress the M S and M F temperatures, so that
some highly alloyed steels must be cooled to below room temperature to obtain fully martensitic structures.
Many steels are quenched in either water or oil to produce adequate cooling rates. While water quenching
produces the fastest cooling rates, it also produces the highest residual stresses and can often cause warpage
and distortion; therefore, higher alloy grades are often used so that a milder oil quench can be applied. The
rate at which a given portion of a steel bar cools from the austenitizing temperature depends on two factors:
(1) the temperature to which the surface of the bar is cooled by the quenching medium, and (2) the rate of
heat flow in the bar itself. Since heat flow in the bar itself is relatively low, the center of a steel bar cools much
more slowly than the surface. Different quench media are denoted by H, which represents the cooling power
of the quenching medium. An ideal quench is one in which there is no resistance to heat transfer from the
bar to the quench media (H=∞), so the surface immediately achieves the temperature of the quench bath,
Representative quench values are shown in Table Table 1.2. While water quenching produces a fairly rapid
cooling rate, it also produces high residual stresses that can lead to cracking and distortion. Since large tem-
perature gradients can cause distortion, it is often necessary to use a milder quench, such as oil or an air
blast. There are three stages of heat removal during quenching in liquids, as illustrated in Fig. 1.16: (1) vapor
blanket stage, (2) nucleate boiling stage, and (3) liquid cooling stage. The vapor blanket stage is character-
ized by the formation of a uniform vapor blanket around the part. The vapor blanket is maintained while the
supply of heat from the interior of the part to the surface exceeds the amount of heat needed to evaporate the
quenchant and maintain the vapor phase. The highest cooling rates occur during the nucleate boiling stage.
During this period, the vapor envelope collapses, and high heat extraction rates are achieved that are associ-
ated with nucleate boiling of the quenchant on the metal surface. Heat is rapidly removed from the surface
as liquid quenchant contacts the metal surface and is vaporized. The liquid cooling stage begins when the
temperature of the metal surface is reduced below the boiling point of the quenching liquid. Below this tem-
14 1. Introduction

Figure 1.14: Continuous cooling transformation diagram for 4340 steel.

perature, boiling stops, and cooling takes place by conduction and convection into the quenchant. Agitation
refers to liquid quenchant movement relative to the part. Agitation is usually obtained by stirring the liquid,
but in some cases, it is obtained by moving the part in the liquid. Agitation causes mechanical disruption of
the vapor blanket and a faster transition to the nucleate boiling stage cooling. However, increasing agitation
usually produces faster cooling rates in all three regions. Non uniform quenching can result if agitation is not
used, due to localized hot spots resulting from uneven heat removal from the metal surface. This can lead to
spotty hardness, increased surface cracking, distortion, and higher residual stresses. Many different media
have been used for quenching, such as water, brine solutions, oils, and synthetic polymer solutions. The ideal
quenchant would have a high initial quenching effect during the vapor phase and boiling range periods but
would cool slowly through the final convection range(liquid cooling phase). Cold water, and especially brine
solutions, have the highest initial quenching speeds, but they also quench very fast at the end of the quench-
ing process, that is, during the convection phase. Thus, they are restricted to quenching simple shapes and
steels of comparatively low hardenability. For other work pieces, they would cause either intolerable degrees
of distortion or warpage and possibly cracking. All quenching oils have considerably lower quenching rates
than water or salt solutions. Moreover, their heat extraction is more uniform and particularly slow at the end
of the cooling cycle, that is, during the convection range. As a result, the dangers of distortion or cracking are
greatly reduced.

1.3.9. Tempering
While as-quenched steel is extremely hard and strong, it is also very brittle. Tempering, a process where the
steel is reheated to an intermediate temperature, is used to increase the ductility and toughness, with some
loss of strength and hardness. During tempering, the highly strained bct structure starts losing carbon to
transformation products, which reduces lattice strains, producing an increase in ductility and a reduction in
strength. Tempering involves reheating the steel to some temperature below A 1 and holding for some period
1.3. Introduction to Thermal Heat Treatment Processes 15

Figure 1.15: Continuous cooling transformation diagram for 4140 steel.

Quench severity, H
Agitation Oil Water Brine
None 0.25–0.30 0.9–1.0 2
Mild 0.30–0.35 1.0–1.1 2.0–2.0
Moderate 0.35–0.40 1.2–1.3 ...
Good 0.40–0.50 1.4–1.5 ...
Strong 0.50–0.80 1.6–2.0 ...
Violent 0.80–1.1 4.0 5.0

Table 1.2: Quench values for several quenching media

Figure 1.16: Three stages of quenching.


16 1. Introduction

of time, usually 1 to 2 h. As in most heat treatment processes, the tempering temperature is much more im-
portant than tempering time. As the tempering temperature is increased, the hardness decreases (Fig. 1.17).
Plain carbon and low-alloy steels can be tempered at lower or higher temperature ranges, depending on the
balance of properties required. Tempering between 150 and 200◦ C will retain much of the hardness and
strength of the quenched martensite and provide a small improvement in ductility and toughness. This treat-
ment can be used for bearings and gears that are subjected to compression loading. Tempering above 430◦ C
significantly improves ductility and toughness but at the expense of hardness and strength. The effect of tem-
pering temperature on the tensile properties of oil-quenched 4340 medium-carbon low-alloy steel is shown
in Fig. 1.18. The tensile strength

Figure 1.17: Effect of carbon content on tempering of plain carbon steels.

has decreased from 1900 MPa at a 200◦ C tempering temperature to approximately 965 MPa at a 650◦ C
tempering temperature. However, the ductility, as measured by total elongation and reduction in area, in-
creases dramatically. The tempering range of 260 to 370◦ C is often avoided because it can result in low impact
toughness (Fig. 1.19) due to temper embrittlement, During tempering, carbon atoms dispersed in martensite
form carbide precipitates of increasing size. Tempering occurs in stages as the steel is heated to higher and
higher temperatures. However, it should be noted that the stages of tempering are somewhat arbitrary and
that they overlap considerably, with reactions continually occurring as the part is heated to higher and higher
temperatures. The three stages of tempering involve: stage 1, the formation of a transition carbide and the
reduction of the carbon content in the martensitic matrix; stage 2, the transformation of any retained austen-
ite to ferrite and cementite; and stage 3, the replacement of the transition carbide and low-carbon martensite
with ferrite and cementite.
Stage 1. The first stage of tempering occurs at temperatures between 95 and 260◦ C. At the beginning
of stage 1 tempering of low-carbon steels, the carbon atoms redistribute themselves to lower-energy sites
such as dislocations. Actually, a large percentage of the redistribution of carbon atoms takes place during
quenching through the temperature range in which martensite forms. For carbon contents of less than ap-
proximately 0.2 wt%, up to 90% of the carbon segregation takes place during quenching. Since the carbon
atoms can reduce their energies more by segregating to dislocation sites than by forming transition carbides,
no transition carbides form in steels with less than approximately 0.2 wt% C. In steels containing more than
1.3. Introduction to Thermal Heat Treatment Processes 17

Figure 1.18: Effects of tempering temperature on 4340 steel.

0.2 wt% C, the initial carbon segregation occurs by precipitation clustering. Very fine particles of transition
carbide then nucleate and grow within the martensite. The carbon content of the martensitic matrix is re-
duced by the formation of transition carbides. The transition carbides can be ǫ carbide (epsilon carbide),
which has the approximate composition formula Fe2 · 4 C and a hexagonal crystal structure, and/or η carbide
(eta carbide), with the formula Fe2 C and an orthorhombic crystal structure. Both epsilon carbide and eta
carbide have substantially higher carbon contents than the cementite Fe3 C that forms later when tempering
is conducted at higher temperatures. When the transition carbides form, the martensite retains some degree
of tetragonality, because it still contains more carbon in solid solution than ferrite does. Therefore, when the
total carbon content is high enough, the first stage of tempering involves the segregation of carbon to various
defects in the microstructure and the conversion of martensite to low-carbon martensite and a transition car-
bide. During stage 1 tempering, there are also changes in the physical properties, such as electrical resistivity,
which can be used to monitor the progress of the changes. However, there is not much of a reduction in the
hardness; in fact, it can increase slightly for steels of medium-to-high-carbon contents.
Stage 2. The second stage of tempering occurs between 205 and 315◦ C. It is not very important in low-
carbon steels, because it involves the conversion of retained austenite to ferrite and cementite. The quantity
of retained austenite in low-carbon steel is not large. However, it can be large in alloy steels, especially those in
which the M F temperature is below room temperature. Thus, stage 2 is important when there are significant
quantities of retained austenite in the martensite. The transformations of retained austenite in 4130 and 4340
steels, which contain approximately 2 and 4 vol% of retained austenite, respectively, are shown in Fig. 1.20.
For tempering times of 1 h, the transformation of retained austenite begins above 205◦ C and is complete by
315◦ C. Note that the percentage of cementite increases as the amount of retained austenite decreases.
Stage 3. The third stage of tempering, which begins at approximately 260◦ C, results in the formation of
ferrite and cementite. An additional transition carbide, known as chi (χ) carbide, with the formula Fe5 C2
and a monoclinic structure, has been identified in high-carbon steels during tempering between 205 and
315◦ C. It is a metastable carbide with a composition intermediate between epsilon carbide and cementite.
As the tempering temperature is increased in the range of 260 to 705◦ C, the transition carbides dissolve and
18 1. Introduction

Figure 1.19: Effects of tempering temperature on impact toughness.

are replaced by cementite, Fe3 C. The diffusion of carbon to cementite causes the martensite matrix to lose
tetragonality and become bcc ferrite as its carbon content drops. The initial shape of cementite is needle
like when tempered between 205 and 315◦ C. In high-carbon martensite, Fe3 C nucleates at twin boundaries,
and in low carbon lath martensite, it nucleates at the martensite lath boundaries as well as at dislocations.
From 400 to 595◦ C, the lath like carbides coalesce to form spheroidite, which reduces the overall surface en-
ergy. From 595 to 704◦ C, spheroidite coarsens even more, with the smaller particles dissolving. The driving
force for coalescence is the reduction of the overall surface energy of the cementite in the ferrite matrix. The
hardness after tempering decreases, an effect that becomes more and more significant as the tempering tem-
perature becomes higher. Eventually, the temperature becomes high enough to cause the dislocation content
to decrease in a manner similar to the changes in the dislocation content of cold-worked metals during recov-
ery processes. At approximately 595◦ C, the dislocation density is substantially reduced. The lath morphology
in lath martensite and the twin boundaries in plate martensite disappear, and the ferrite grains grow and
become equiaxed. These morphological changes in the ferrite grains are similar to those that occur dur-
ing recrystallizationin cold-worked ferrite. Secondary hardening of certain alloy steels can also occur during
stage 3 tempering, although secondary hardening is sometimes referred to as stage 4 tempering. Except for
stage 1, the hardness and strength of the steel decreases during tempering, while the ductility and toughness
increase. Although these various stages offer a convenient way to classify the different tempering reactions,
it is important to recognize that the temperatures are merely guidelines. The stages overlap because the
dominant type of transformation and the reaction kinetics depend on temperature and time, both of which
change continuously. Alloying elements slow the rate of tempering and lessen the decrease in hardness. The
tempering process can be retarded by the addition of alloying elements such as vanadium, molybdenum,
manganese, chromium, and silicon. These larger substitutional alloying elements diffuse slower than the in-
terstitial carbon, and some of them form carbides that are harder than cementite. In some high-carbon steels
with significant alloying additions, some retained austenite will convert to martensite on cooling from the
tempering operation. Since this adversely affects the toughness, such steels are often double or even triple
tempered. Some alloy steels containing molybdenum, chromium, and tungsten exhibit secondary hardening
during tempering. These elements lead to precipitation hardening during tempering. The cementite-rich
1.3. Introduction to Thermal Heat Treatment Processes 19

Figure 1.20: Transformation of retained austenite in 4130 and 4340 steels.

carbides are replaced with a new fine array of alloy carbides. Rapidly diffusing carbon enables relatively pure
cementite to form first because the nucleation and growth of the alloy carbides is paced by substitutional
diffusion. When the dense array of fine alloy carbides forms, gliding dislocations meet a higher density of ob-
stacles to slip, and the hardness increases. The particular temperature at which secondary hardening occurs
depends on the specific carbide that forms and increases as the stability of the carbide increases. Secondary
hardening is quite important in certain tool steels, where it helps to retain hardness at elevated temperatures.
Full hardening, followed by tempering, usually results in the best combination of properties. However, in
practice, some pearlite and/or bainite may be mixed with the martensite if the local cooling rate is less than
the critical rate necessary to avoid the nose of the CCT diagram. Retained austenite may also be present if the
M F temperature is below room temperature. In both cases, the properties will not be as good as those with a
fully hardened and tempered martensite matrix.
2
TEMPCORE©, A Low Cost High Strength
Rebar.
In this part the process technique and the general characteristics of TEMPCORE reinforcing steel are re-
viewed. The superior properties in strength, ductility, formability and weldability of this steel as well as the
economic advantages of this technique are compared with those of conventional types of reinforcing steels.
The following Fig. 2.1 represents the normal microstructure of QST rebar with distinct layers and the extra
ductility gained from Quenching and tempering process.

Figure 2.1: cross section of TEMPCORE specimen

2.1. A Breif Historical Background


In the late 1960s and the early 1970s, reinforcing steel users were oriented towards high strength reinforcing
steels with adequate ductility, good weldability, and thus manufacturers were to develop methods leading to
superior bars with relatively low production cost. Conventional methods by hot rolling and cold-working of
carbon steels were out of consideration because of the resulting properties, while micro-alloying methods
promised a certain degree of improvement at an extra cost. Heat treatment has come to the foreground and
in-line heat treatment, by taking advantage of the rolling heat, appeared to be the economical avenue towards
the new product.

21
22 2. TEMPCORE ➞, A Low Cost High Strength Rebar.
As early as 1967 in the former USSR (Union of Soviet Socialist Republic) an attempt was made to produce
high strength reinforcing steel bars by this method (Chemenko et al. 1987). Another in-line heat treatment
technique, named TEMPRIMAR, was developed in the former Federal Republic of Germany (Vlad 1985; Klaus
and Vlad 1985). But amongst the in-line heat treatment methods, the TEMPCORE process became the most
successful and the most popular. This latter method originates from the laboratories of Centre de Recherches
Metallurgiques (C.R.M), and in 1972, TEMPCORE process was patented in Belgium (Belgian Patent 790.867
of 31-10-1972).
In 1974 the first industrial-scale trials were carried out in Esch-Schifflange Division of S. A. ARBED and this
was the beginning of industrial scale manufacturing of TEMPCORE reinforcing steel. In 1975, Economopou-
los et al. (1975) published a paper entitled "Application of the TEMPCORE process to the fabrication of high
yield strength concrete-reinforcing bars" in Metallurgical Reports C.R.M., in which the principles of the TEM-
PCORE process is described together with the laboratory and industrial trial results. In the same periodical,
Defoumy and Bragard (1977), reported details on the weldability of the steel bars produced by TEMPCORE
process. Rehm and Russwurm (1977) systematically assessed the mechanical properties of this type of re-
inforcing steel. It was demonstrated that the process is capable to produce 500 to 600 MPa grade deformed
reinforcing bars with super ductility and excellent weldability. These steels fulfilled the requirements pre-
sented in DIN 4099 and DIN 488. It was also proved that the process itself is quite stable. An analysis on the
process was published by Simon et al. (1984b), indicating the economic advantages offered, and Kilmore and
Warrett (1984) reported that a further improvement can be achieved by microalloying.
Recently the TEMPCORE process is used in European countries, USA and Australia, and all the other
countries with annual production reach millions tones. TEMPCORE is the best process for the production
of high quality rebars. Costly alloying elements (Vanadium, Niobium) are replaced by low-cost water. It al-
lows obtaining from low C-Mn steel concrete reinforcing bars with high mechanical properties (World-Class
Standards) excellent Weldability and superior ductility and bendability. The process lead to an increase of the
yield strength of 150 to 230 MPa, depending on the cooling intensity, which meaning reducing the alloying
additions and cost saving. Such process design guarantees excellent process controllability (giving consis-
tent product properties) and high flexibility (allowing to produce several grades with the same chemistry).
Equipment manufacturing is then often carried out locally at low cost.
TEMPCORE is a relatively new type of high strength reinforcing bar with excellent ductility and weldabil-
ity. The name is a Trademark and is made up from the two major events taking place during the production
technology. The process involves a quenching immediately after the bar emerges. From the last hot rolling
strand and a self-tempering, by the retained heat from the core, hence the core tempered martensitic prod-
uct, that is: TEMPCORE. The bar in the finished state is made up by tempered martensitic hardened layer and
ferrite-pearlite core with an intermediate hardened layer of a mixture of bainite and ferrite in between. Such
a composite steel structure offers many advantages over the other types of reinforcing steels, in strength, in
ductility, weldability and production cost and therefore, since this technology was invented a considerable
expansion took place in the use of TEMPCORE reinforcing steel. The process is a Belgian invention by the
Centre de Recherches Metallurgiques (C.R.M) and the first production tonnage date back to 1974. Although
the production output increased significantly in the world generally, and in spite of the sophisticated metal-
lurgical composition of the steel itself, literature survey shows a lack of published data on this type of steel
which concerning production of QST steel rebars with high production rate close to the same of hot-rolling
production, almost all the previous work concentrated on evaluation of QST rebars produced by QTB regard-
less of the speed of the rolling line with introducing the Quenching box. The aim of production with high
production rate not in the scope of most producers in spite of achieving high productivity beside high quality
are the main target for any steel rebars manufacture.

2.2. The Tempcore Process


TEMPCORE Process has been developed in the early seventies by C.R.M., Belgium in order to manufacture
high yield strength weldable concrete reinforcing bars from mild steel, without V or Nb additions. It consists
in subjecting the hot rolled steel to an in-line heat treatment in 3 successive stages. Tempcore is the Best Pro-
cess for the Production of High Quality Rebars: Costly alloying elements (Vanadium, Niobium) are replaced
by low-cost water, in order to obtain from low C-Mn steel, concrete reinforcing bars with:

1. High mechanical properties.

2. Excellent weldability
2.2. The Tempcore Process 23

3. Excellent ductility and bendability.

TEMPCORE process, which produces high strength reinforcing bars from plain low carbon steel, is basically
an in-line heat treatment process. It consists of quenching and self-tempering from which the name is de-
rived, i.e., a quenched surface layer is Tempered by the heat dissipating from the CORE. As it is illustrated in
Fig. 2.2, steel billets are heated to approximately 1150◦ C in reheating furnaces And rolled through a sequence
of rolling strands which progressively reduce the billets to the final size and shape of reinforcing bars. On
leaving the last strand, a controlled cooling is applied in such a way that the bar undergoes three stage met-
allurgical transformations as it is illustrated by a typical CCT diagram in Fig. 2.2 below. On leaving the last

Figure 2.2: Transformation in a TEMPCORE Process

rolling strand, the bar passes through a water quenching chamber, an intensive cooling forces the tempera-
ture within a certain depth from the surface of the bar down to below the martensite transformation starting
point (M s ). Martensitic structure is formed within this quenched layer while the core remains austenite and
a large temperature gradient is established in the cross section of the bar. During the second stage, when
the bar leaves the quenching chamber and is exposed to air cooling, the heat retained in the core dissipates
through the hardened layer reheating the quenched martensite, a self-tempering is experienced. Tempered
martensite is the result from a "tempering temperature". At the same time, the layer adjacent to the hard-
ened layer starts to transform to bainite and an intermediate and intermediately hardened layer is formed.
The core still remains austenite. During the third stage, the remaining austenite in the core undergoes an
essential quasi-isothermal transformation forming ferrite-pearlite structure. Consequently, a TEMPCORE
reinforcing steel bar is essentially a composite material consisting of concentrically disposed hard outer layer
and soft core with an intermediate and intermediately hardened layer. With relatively low carbon content,
TEMPCORE reinforcing steel provides high strength, super-ductility and excellent weldability amongst other
advantages.

2.2.1. Theory of the TEMPCORE Process


The purpose of this section is to analyze the fundamental physical phenomena underlying the TEMPCORE
process and to discuss the effect of the more important independent variables on the final result of the TEM-
PCORE heat treatment, namely the mechanical properties of the treated bars. Principle of the Process TEM-
PCORE is a direct quenching and self-tempering process from the heat of rolling. The present description
is exclusively devoted to the heat treatment of hot rolled concrete reinforcing bars. According to the TEMP-
24 2. TEMPCORE ➞, A Low Cost High Strength Rebar.

Figure 2.3: Schematic of Quenching and self-tempering process

CORE process the reinforcing bar from the rolling mill is submitted to a special heat treating cycle involving
three stages as shown in Fig. 2.3 :

First Stage: The first stage is a fast cooling operation applied to the bar as it leaves the last finishing stand. The
efficiency of the cooling facility used in this first stage must be such that the cooling rate down to a certain
depth below the skin is higher than the critical rate for martensitic quenching. At the conclusion of this oper-
ation, the bar has an austenite core surrounded by a layer composed of a mixture of austenite and martensite,
with the martensite content decreasing from the skin towards the inner. The aimed duration of the first stage
depends on the desired thickness of the martensitic layer.

Second Stage: During the second stage, the bar leaves the area of drastic cooling and is exposed to air. The
heat transfer coefficient of the environment being very small, and the temperature gradient within the bar’s
cross-section being very large, the core reheats the quenched surface layer by conduction. As a result, the
martensite formed during the first stage is subjected to self-tempering which, as will be shown below, en-
sures adequate ductility while maintaining a high yield strength level. As soon as this quenching operation is
stopped, the surface layer is Tempered by using the residual heat left in the CORE of the bar (self-tempering
of the martensite layer); hence the name TEMPCORE. The process ensures uniform formation of the wall-
thickness of each layer all throughout the Rebar.
The second stage is conventionally considered as terminated when the temperature at the surface of the
bar passes through a maximum, called “Tempering Temperature”. The duration of the second stage varies
greatly with the bar diameter and the cooling conditions applied during the first stage. In the course of the
second stage, the untransformed austenite in the surface layer transforms to bainite, whereas the core re-
mains austenitic. On the other hand, the austenite subjacent to the tempered martensite layer can start to
transform to bainite, depending on the Steel’s composition and the cooling conditions.

Third Stage: The third stage occurs as the bar lies on the cooling bed. It consists of a quasi-isothermal trans-
formation of the remaining austenite. The product of this transformation is a mixture of either ferrite and
pearlite or ferrite, pearlite and bainite, depending on several factors:

1. The steel composition

2. The bar diameter

3. The finishing temperature of the rolling sequence

4. The efficiency and duration of cooling during the first stage

The physical phenomena involved in the above mentioned three stages of the Tempcore process as shown in
Fig. 2.4 can be divided in three categories:

1. Heat exchanges between the surface of the bar and the surroundings

2. Heat conduction in the bar

3. Metallurgical phenomena such as allotropic transformations.


2.2. The Tempcore Process 25

In the following sections, the effect of the governing parameters and the interrelations of the above phenom-
ena will be discussed.

Figure 2.4: Schematic diagram of TEMPCORE process

TEMPCORE is the best process for the production of high quality rebars. Costly alloying elements (Vana-
dium, Niobium) are replaced by low-cost water. It allows obtaining from low C-Mn steel concrete reinforcing
bars with high mechanical properties (World-Class Standards) excellent Weldability and superior ductility
and bendability. The process lead to an increase of the yield strength of 150 to 230 MPa, depending on the
cooling intensity, which meaning reducing the alloying additions and cost saving. Fig. 2.5 illustrates the im-
proved in ductility, weldability and bendability gained from TEMPCORE process even when achieving the
same mechanical properties Fig. 2.5 shows weldability and ductility properties of TEMPCORE rebars Such

Figure 2.5: shows weldability and ductility properties of TEMPCORE rebars

process design guarantees excellent process controllability (giving consistent product properties) and high
flexibility (allowing to produce several grades with the same chemistry). Equipment manufacturing is then
often carried-out locally at low cost.
26 2. TEMPCORE ➞, A Low Cost High Strength Rebar.
2.2.2. Fundamental Mechanisms of Tempcore Hardening
Fig. 2.6 shows a schematic representation of the fundamental mechanisms of Tempcore hardening and their
interrelations (CRM, 1985). The involved phenomena have been divided as stated previously in three classes:
surface heat exchanges, heat conduction and physical metallurgy (vertical dotted lines in Fig. 2.6. On the top
of the Fig. 2.6 are represented, inside circles, the governing parameters of the process. Among them, only
three can be considered as independent control variables from the point of view of the rolling mill operator,
namely: the water flow rate, the quenching time and the finishing temperature. In Tempcore treated bars,
the microstructure and properties vary continuously from the surface to the axis of the bar. Nevertheless, it
is possible, as a satisfactory approximation to consider the Tempcore bars as composed of two distinct parts:
a surface layer that for the sake of simplicity, we shall name “Tempered Martensite” or “Mertensite” and the
central part or “Core” composed of ferrite and carbides.

Figure 2.6: fundamental mechanisms of Tempcore hardening and their interrelations

The overall macroscopic properties of the bar, particularly tensile properties, depend on three factors:

1. The volume percentage of martensite.

2. The tensile properties of martensite.

3. The tensile properties of the core ferrite-carbides structure.

The volume percentage of martensite depends on the starting temperature of martensite transformation
(M s point), itself a function of composition and the temperature field in the cross-section of the bar leaving
the quenching device. It is conventionally considered that the limit of the martensite layer corresponds to a
circular cylinder of the same axis as the bar and for which the minimum temperature is equal to the M s point
The tensile properties of the martensitic layer depend on the chemical composition and on the “Tempering
Temperature”. This tempering temperature is itself a function of the thermal field in the cross-section of the
bar leaving the Tempcore cooling device. The mechanical properties of the core depend on two groups of
parameters:

1. The chemical composition through CCT diagrams and microstructure-properties relations

2. The cooling conditions during the quenching and the subsequent cooling stages.

It results from the above analysis that, for a given chemical composition, the main factor determining the
mechanical properties of the bars is the thermal field resulting from the quenching stage. This thermal field
is governed by the well-known laws of conduction of heat in solids For a given bar diameter, the thermal field
can be modified by changing the temperature of the bar entering the quenching device, the duration of the
quenching stage and the heat transfer coefficient between the bar surface and the cooling water. The heat
2.2. The Tempcore Process 27

transfer coefficient is the key factor of the TEMPCORE process; it is expressed as a function of the surface
temperature of the bar, function depending mainly on the design of the cooling device and the cooling water
flow rate and temperature. The above analysis shows that the hardening mechanism of TEMPCORE bars is an
intricate one; nevertheless, we shall show that, despite this fact, it is possible to find relatively simple relations
describing the more important phenomena.

2.2.3. Surface Heat Exchanges


As far as surface heat exchanges are concerned, a TEMPCORE heat treating line can be divided in three sec-
tions: the drastic water quenching section, the air cooling section during the travel of the bar between the exit
of the quenching line and the cooling bed, the air cooling of parallel bars lying on the cooling bed. In order to
describe the cooling on each of the above sections in a practical sense (i.e. disregarding the complicated fun-
damental mechanisms of cooling), we use the very convenient parameter called “Heat Transfer Coefficient”
(α) and defined as follows:

Heat F l ux Densi t y (ϕ) = α(T s –Tm )

Where: T s represents the surface temperature of the bar and Tm the bulk temperature of the surrounding
cooling medium.
In general, the heat transfer coefficient is very strongly dependent on the surface temperature and on the
specific parameters of the cooling device considered. Fig. 2.7 shows a typical heat transfer coefficient-surface
temperature curve. During the travel of the bar between the exit of the quenching line and the cooling bed,
cooling results from a combination of radiation and forced air convection. The heat transfer coefficient is a
function of the surface temperature of the bar and of the relative speed between the bar and the surrounding
air. The heat transfer coefficient on the cooling bed results from a combined radiation and natural convection
cooling. Due to the view factor effect, the heat transfer coefficient depends not only on the surface temper-
ature of the bar but also on the bar diameter and the distance between the notches of the cooling bed. Ta-
ble 2.1 shows the changing heat, capacity and heat conduction coefficient as a function of internal structure
and temperature. Changing internal structure and temperature change heat capacity and heat conduction
coefficient.

Temperature Structure 0◦ C 300◦ C 600◦ C 9000◦ C


Conduction γ 0.004 0.005 0.006 0.007
Coefficient K α,p,b,m 0.013 0.0105 0.009 0.007
(cal./mm.sec◦ C)
Heat Capacity γ 0.00076 0.00086 0.00095 0.00105
ρC P α,p,b,m 0.00084 0.00092 0.00102 0.00110
(cal./mm3◦ C)

Table 2.1: changing heat capacity and heat conduction coefficient as a function of internal
structure and temperature. Changing internal structure and temperature change heat
capacity and heat conduction coefficient.

2.2.4. Heat Conduction in the Bar


It has been shown that, for a given chemical composition, the final tensile properties of a TEMPCORE treated
bar depend exclusively on the thermal field in the bar at the moment it leaves the quenching devices. In
the following paragraphs, the dependence of the thermal field on the different process parameters will be
discussed in more details. As a first step, the problem of heat conduction in the bar has to be solved; its
solutions require:
• The definition of the boundary conditions. The knowledge of the heat transfer coefficients presented
in the previous paragraph constitutes a satisfactory definition of these boundary conditions.

• The knowledge of the thermal properties of the steel. Temperature dependent properties gathered from
technical literature are used.

• The numerical solutions of the Fourier differential equation and its implementation in a computer
program.
28 2. TEMPCORE ➞, A Low Cost High Strength Rebar.

Figure 2.7: typical heat transfer coefficient-surface temperature curve.

The solution of the heat conduction problem enables the calculation of the temperature at any moment and
any point of the cross-section of the bar. Having thus determined the thermal field in the bar, it is possi-
ble to calculate the two important parameters from a metallurgical point of view: the volume percentage of
martensite (p m ) and the “theoretical tempering temperature” (T ∗ ) defined below. The following sections will
be devoted to the study of the effect of quenching conditions on (p m ) and (T ∗ ).

2.2.5. Volume Percentage of Martensite (p m ):


Fig. 2.8 shows the temperature distribution in the cross-section of a 25 mm diameter bar after 1, 1.5 and 2
seconds quenching. It is interesting to observe the very steep temperature gradient near the surface; this is
an important characteristic of the TEMPCORE quenching equipment because it enables to form a relatively
thick layer of martensite and, in the same time, to leave enough heat in the core for the self-tempering op-
eration. For each of the three quenching times, Fig. 2.8 shows the radius limiting the martensitic layer (r m ).
2.2. The Tempcore Process 29

Figure 2.8: cross section of TEMPCORE specimen

The starting temperature of the martensitic transformation (M s ) is a function of chemical composition. The
following formula is sufficient for our study:

M s = −361(C %) − 39(Mn%) + 500

If(d) represents the equivalent diameter of the bar, the volume percentage of martensite is expressed by the
relation:
2r m 2
· µ ¶ ¸
P m = 100 1 −
d
In this example p m = 30.8 %, 24.3 % and 16.8 % for the quenching times of 2 sec., 1.5 sec. and 1 sec. re-
spectively. For a given cooling device, water flow rate per unit length and water temperature, and for a fixed
chemical composition, the percentage of martensite is a function of the quenching time, the diameter of the
bar and its entry temperature in the quenching device. The relation p m - quenching time as shown in Fig. 2.9
can be considered with a very good approximation as linear in the useful range of (p m ) (CRM, 1985). The
slope of this curve increases sharply with decreasing diameters. We shall show later that this rapid increase of
the slope does not create any problem as far as the controllability of the process is concerned. Fig. 2.9 shows
also (compare lines c and d) the effect of the initial quenching temperature: the slope of the “p m - time” line
is not affected but (p m ) significantly decreases when (To ) increases. This fact has to be kept in mind when
discussing the effect of thermomechanical treatment.

2.2.6. Tempering Temperature


The tempering temperature has been defined as being the maximum temperature reached at the surface of
the bar during the self-tempering stage. This parameter (Tr ) is important in practice for two reasons:

1. it corresponds to the minimum tempering temperature of the martensitic surface layer.

2. it can be measured directly by a radiation pyrometer and be used for controlling the process.
30 2. TEMPCORE ➞, A Low Cost High Strength Rebar.

Figure 2.9: Effect of Quenching time, bar diameter and initial temperature on the Martensite content

From the point of view of the theoretical analysis of the process, we shall use another parameter called “the-
oretical tempering temperature”, symbolized by (T ∗ ) and corresponding to the following definition: let it be
the enthalpy per unit length of the bar at the exit of the quenching device; (T ∗ ) is the uniform temperature
of a bar in all respects identical to the considered one and having the same enthalpy per unit length. The in-
terest of (T ∗ ) is that its value is exclusively linked to the temperature field at the exit of the quenching device:
the phenomena occurring during the air cooling phase have not to be taken into account. The difference (T ∗
- Tr ) depends on the heat transfer coefficient in the forced air cooling section and on the time necessary for
the surface of the bar to reach (Tr ). As this time increases with the bar diameter, so does the difference (T ∗
- Tr ), this kind of behaviour is shown in Fig. 2.10. With the above remarks in mind, one can now discuss the
effect on (T ∗ ) of the bar diameter (d), the initial temperature (To ) and the quenching time (t). Fig. 2.11 shows
that the relation T ∗ (t ) is linear and that the slope is strongly dependent on the diameter (CRM, 1985). The
initial temperature has only a translation effect on T ∗ (t ). The relation between the above parameters can be
expressed analytically by:

T ∗ = −Ad −β t + f (T0 , d ) where α and β are constants

2.2.7. Control Function of the Process


As (Tr ) and (p m ) are the main controlling factors of the tensile properties, it is interesting to establish a re-
lation between them. Fig. 2.12 shows that the relation ‘’p m − (Tr )” is linear and that the slope is approxi-
mately independent on the bar diameter (CRM, 1985). With increasing diameters, the “p m − (Tr )” line un-
dergoes a parallel translation towards increasing tempering temperatures. For practical applications, a single
“p m −(Tr )” relation can be considered for diameters between 16 mm and 40 mm. The effect of initial quench-
ing temperature on the relation “p m − (Tr )” is shown in Fig. 2.12 (CRM, 1985). In practice and for normal
operating conditions, the variation of finishing temperature for the same diameter does not exceed +50◦ C.
The same maximum range of variation can be foreseen for (To ). In that case, we can neglect, for practical
purposes, the influence of entry temperature on the relation “p m − (Tr )”. From the above analysis, one can
conclude that to each pair (d − To ) corresponds a relation (p m − Tr ) that will be used to control the process.
For 16 ≤ d ≤ 40 mm, a single relation can be defined.
2.2. The Tempcore Process 31

Figure 2.10: Effect of different parameters on T ∗

2.2.8. Mechanical Properties


In the previous sections, the effect of the independent variables of the process on two important parameters
has been discussed: the volume percentage of martensite (p m ) and tempering temperature (T ∗ or Tr ). In
this section, (p m ) and (Tr ) will be considered as the independent variables and their effects on mechanical
properties will be discussed. As metallurgical phenomena are involved, an additional important variable will
be introduced: chemical composition. For the steel grades treated by the Tempcore process, carbon and
manganese (in some cases silicon) contents are the main components characterizing the composition from
a practical point of view. In two separate sections the properties of the tempered martensite and those of the
core of the treated bars will be discussed. In a third section, the combination of the two individual zones of
the bar leading to the final properties, will be discussed. Tempered Martensite The properties of tempered
martensite depend exclusively on its composition and on the tempering temperature. In order to determine
these relations in this case of Tempcore steels, specimens of different compositions have been quenched in
still water and submitted to a subsequent tempering treatment (CRM, 1985). The following relation gives the
yield strength (expressed in MPa) of the martensitic layer with a precision sufficient for practical uses:

Ym = −1.75Tr + 1781
32 2. TEMPCORE ➞, A Low Cost High Strength Rebar.

Figure 2.11: effect of diameter, initial temperature and quenching time in the theoritical tempering temperature

Figure 2.12: the effect of bar diameter on the (p m − Tr ) relation


2.2. The Tempcore Process 33

Figure 2.13: the effect of initial temperature on the (p m − Tr ) relation

2.2.9. Central Region of the Bar


One calls “central region of the bar” or “core” the part of the bar in which the temperature at the exit of
the quenching line is higher than the M s temperature. It is obvious that at each point of the cross section
corresponding to the core, the cooling curve and consequently, the properties are different. This situation

Figure 2.14: CCT Diagram at different points along the cross section of the rebar
34 2. TEMPCORE ➞, A Low Cost High Strength Rebar.
Table 2.2: Isothermal treatments in lead bath

Composition (%) Yc (Mpa) for Tc (◦C )


C Mn Si N2 550 570 600 625 650
0.139 0.64 0.037 0.014 365 354 340 334 318
0.120 0.59 0.025 0.012 340 330 317 309 298
0.211 0.79 0.005 0.001 371 357 354 351 349
0.112 0.93 0.026 0.004 335 317 310 304 301
0.199 0.875 0.005 0.009 380 362 356 352 347
0.130 1.213 0.03 0.013 402 384 366 350 330
0.148 0.955 0.024 0.004 337 324 313 304 294

being too complicated to be handled by simple analysis, it has been simplified drastically in this section.
The observation of Fig. 2.14 suggests a stylization of the mean cooling curve of the core (CRM, 1985). The
cooling of the core can be assimilated roughly to an isothermal transformation at temperature Tc . In order
to determine the relation between the yield strength of the core (Yc ) and (Tc ) and the composition, some
treated specimens of the steels are listed in Table 2.3 by austenizing and quenching in a lead bath at different
temperatures (Tc ) (CRM, 1985). A regression analysis of the results given in Table 2.3 leads to the following
formula:
Yc = −0.406Tc + 357(C %) + 38.7(Mn%) + 495

The observation of the previous Table 2.3 shows that, at first approximation, the value of (Tc ) in the above
formula can be taken equal to (Tr ).

2.2.10. Overall Yield Strength of the Tempcore Bar


Having determined the yield strength of the two regions into which the cross-section of a Tempcore bar has
been artificially divided, one seeks a way to combine (Ym ) and (Yc ) in order to foresee the overall yield strength
of the bar. Let us think of a Tempcore bar as being divided in a series of concentric tubes, the mechanical
properties varying from one tube to the other. As there is a continuity between a tube and its neighbors the
problem consisting to calculate the overall properties knowing the individual properties of the tubes, is a
complicated one. Two important simplifications, justified by experimental evidence, will be introduced:

1. additivity of the tensile forces. This hypothesis has been verified with a reasonable precision by tensile
tests on Tempcore bars machined at progressively smaller diameters and also on tubes machined by
boring bars to progressively higher diameters.

2. the continuous variation of properties is neglected and only two zones are considered: a surface layer
having a yield strength (Ym ) and occupying (p m )% of the volume and a core zone with a yield strength
(Yc ) and occupying (1 - p m ) % of the volume.

It is easy to demonstrate that under the above conditions:

Y = P m Y M + (1 − P m )Yc

Finally, the Tempcore process is described by the following relations:

1. p m = f (t , d , T0 , M s )

2. Tr = f (t , d , To )

3. M s = −361C − 39Mn + 500

4. Ym = 1.75Tr + 1781.5

5. Yc = −0.406Tr + 357C + 38.7Mn + 495

6. Y = P m (Ym − Yc ) + Yc
2.3. Characteristics of Tempcore Reinforcing Steels 35

Table 2.3: Specified chemical composition for Grade B400B-R (400


MPa) reinforcing bars(After EZDK)

Production
Size C% Mn% V%
process
Hot-rolling 0.35 1.00 0.020
D16
QTB 0.23 0.55 —
Hot-rolling 0.35 1.00 0.020
D18
QTB 0.23 0.55 —
Hot-rolling 0.35 1.20 0.040
D22
QTB 0.23 0.55 —
Hot-rolling 0.35 1.20 0.040
D25
QTB 0.28 0.65 —

2.2.11. Practical Conclusions of the Theoretical Study


The mathematical model represented by previous six relations will now be used in order to derive some in-
teresting practical conclusions. Let us consider a typical Tempcore steel containing 0.18 % C and 0.8 % Mn.
The relations (1) to (6) can be combined in the following formula:

Y = P m (1191 − 1.344Tr ) − 0.406Tr + 590

Now we can establish a relation between the yield strength of the bar and the tempering temperature.

2.3. Characteristics of Tempcore Reinforcing Steels


2.3.1. Type of steel
TEMPCORE reinforcing steels are basically plain low carbon steels specified for yield strength, ductility, car-
bon or carbon equivalent and yield to tensile ratio. The maximum and minimum specified carbon content
intends to ensure weldability and hardenability. With too low carbon content hardenability of the steel will
not be sufficient and thus more sever quenching is required affecting rolling mill design, e.g., speed of rolling
mill, length and efficiency of cooling chamber. Semi-killed carbon steel with 0.13% - 0.28% carbon content
and the carbon equivalent (CE) < 0.50% has been proved to be the best balance to satisfy the above considera-
tions. The chemical composition of the steel grades for different sizes is shown in the next Table 2.3 indicating
clearly that TEMPCORE reinforcing steel has the leanest chemistry:

2.3.2. Metallurgical phases and microstructures


The microstructure is usually fine due to a relative fast cooling in the core and to the thermomechanical treat-
ment involved in TEMPCORE process, e.g., polygonal ferrite grains in the core region can be as small as 8 µm
diameter and even 3 µm diameter when lower tempering temperature is applied (Kilmore and Barrett 1984;
Kilmore et al. 1985). However, coarse conglomerate of pseudo-eutectoid and Widmanstatten ferrite in the
core are also possible outcome of the process. Vlad (1985) in his article on "A comparison between the TEM-
PRIMAR and TEMPCORE processes for production of high strength rebars" described that TEMPCORE pro-
cess has a tendency to form Widmanstatten ferrite due to the "inherent higher equalization temperatures".
It is possible that high finishing temperature and perhaps also insufficient rolling deformation are the major
reasons for forming this type of microstructure. High finishing temperature and insufficient rolling deforma-
tion results in large austenitic grains at the end of rolling, and thus coarse martensite and bainite develops in
the hardened layer and in the intermediate hardened layer during the subsequent quenching. Large austen-
ite grain size in the core prevents the impingement of grain boundary ferrite, thus allowing Widmanstatten
ferrite to grow (Honeycombe and Han cook 1981).

2.3.3. Effects of process parameters and steel composition


Naturally if the martensite layer is thicker the retained heat is less and thus the tempering is more modest so
that the bar will exhibit higher yield strength and lower elongation. The process parameters and steel com-
positions play part in the final properties. Longer quenching time, lower finishing temperature and higher
intensity of quenching result in thicker martensitic layer and lower tempering temperature. Higher carbon
36 2. TEMPCORE ➞, A Low Cost High Strength Rebar.
and manganese content increases the hardenability of the steel, and therefore more martensite is formed.
Additionally, the strength of tempered martensite increases as the carbon content increases. Simon (1990),
based on the experiences accumulated during the commissioning of more than 25 TEMPCORE installations,
developed a model which describes the relationship between yield strength and all influencing parameters.
The model was originally used for the design of installations, and it is rewritten as:

τ · qd · F t
Y Sc = and T S = K 2 · C α · Mn β · T Y γ · φδ
K 1 · φa · T0b

Where:

• YS = yield strength, MPa t = quenching time, second q = linear water flow rate, mm3/hper meter of line

• F = filling coefficient, i.e. , φ2 /ID2 where ID = internal diameter of cooling nozzle

• φ = bar diameter, mm

• To = entry temperature of the bar, ◦C

• C = carbon content of the steel, %

• Mn = manganese content of the steel, %

• TS = tensile strength, MPa

• K1,K2,a,b,c,d,e,α,β,γ and δ are constants

Although tempering temperature does not appear in this model directly, the finishing temperature, quench-
ing time, bar diameter and water flow rate relate to it quite strongly. Elongation of TEMPCORE reinforcing
steel has a virtually linear correspondence with the yield strength. From the work of Economopoulos et al.
(1975) it is seen that for a 0.24%C, 1.12%Mn steel, 30% elongation corresponds to yield strength about 420
MPa, and only 17% elongation is obtained when the process is changed to give 680 MPa yield strength.

2.3.4. Tensile properties


The TEMPCORE process can increase the yield stress by 150 to 200 MPa for a given composition (Economopou-
los 1981; Kilmore and Barrett 1984) without losing much elongation. The tensile properties of the bars depend
on the process parameters and steel composition. The range of typical yield strength of TEMPCORE reinforc-
ing steel is between 410 to 550 MPa and elongation on a 5d gage length is 30% down to 25% in the same order.
Typical stress strain curves from Rehm and Russwurm (1977) and Defoumy and Bragard (1977) is show in
Fig. 2.15.
The features are summarized below:

• elastic modulus is 200,000 MPa

• the bar has marked yield point and a Luders type of yield and therefore the 0.01% proof stress (σ0.01 coincides
with 0.2% proof stress (σ0.2 )

• the ratio of yield stress to tensile strength is approximately 0.85

• The bar has large elongation (25% to 30%), large Luders strain and large uniform strain.

TEMPCORE reinforcing steels have two major features when the tensile properties are compared with
those of conventional bars:

• Higher ratio of yield strength to tensile strength.

• Larger elongation.

• Formability
2.3. Characteristics of Tempcore Reinforcing Steels 37

Figure 2.15: Stress strain curve of TEMPCORE reinforcing steel (after Rehm and Russwurm 1977)

Another remarkable property of TEMPCORE reinforcing steel is that it has excellent bending and rebending
properties. Despite the hardened outside layer, minimum bend diameter for a 180◦ single bend is specified
as 1d for 12 mm to 28 diameter bars and 2d for 32 mm and 36 mm diameter bars (BHP 1991). According to
BHP (1982 a & b) and Economopoulos et al.(1975), the 20 and 28 mm diameter bars can even be bent without
mandrel. This is far smaller, that is better, than any specification requirement. The bars can also withstand
all the bending and re-bending tests after aging, satisfying the standard requirements Bending operation
requires less energy when compared with other types of bars due to the low tensile strength to yield strength
ratio, 1.18 versus 1.5 for hot rolled bars. It is estimated that 10% to 20 % energy is saved in bending TEMPCORE
reinforcing steel (BHP 1982a).

Figure 2.16: Bending of 20mm TEMPCORE reinforcing steel (after BHP 1982)

2.3.5. Weldability
Weldability of steel is very sensitive to the chemical composition, especially to carbon content and carbon
equivalent (C.E). In welding, equivalent carbon content (C.E) is used to understand how the different alloying
elements affect hardness of the steel being welded. This is then directly related to hydrogen induced cold
cracking, which is the most common weld defect for steel, thus it is most commonly used to determine weld-
ability. Higher concentrations of carbon and other alloying elements such as manganese, chromium, silicon,
molybdenum, vanadium, copper, and nickel tend to increase hardness and decrease weldability. Each of
these elements tends to influence the hardness and weldability of the steel to different magnitudes, however,
making a method of comparison necessary to judge the difference in hardness between two alloys made of
different alloying elements.
There are two commonly used formulas for calculating the equivalent carbon content. One is from the
American Welding Society (AWS) and recommended for structural steels and the other is the formula based
38 2. TEMPCORE ➞, A Low Cost High Strength Rebar.
on the International Institute of Welding (IIW). The AWS states that for equivalent carbon content above
0.40% there is a potential for cracking in the heat-affected zone (HAZ) on flame cut edges and welds. However,
structural engineering standards rarely use CE, but rather limit the maximum percentage of certain alloying
elements. This practice started before the C.E concept existed, so just continues to be used. This has led to
issues because certain high strength steels are now being used that have a C.E higher than 0.50% that have
brittle failures.
%Mn + %Si %C r + %Mo + %V %Cu + %N i
µ ¶ µ ¶ µ ¶
C E = %C + + +
6 5 15
The other and most popular formula is the Dearden and O’Neill formula, which was adopted by IIW in 1967.
This formula has been found suitable for predicting hardenability in a large range of commonly used plain
carbon and carbon-manganese steels, but not to micro alloyed high-strength low-alloy steels or low-alloy
Cr-Mo steels. The formula is defined as follows:
%Mn %C r + %Mo + %V %Cu + %N i
µ ¶ µ ¶
C E = %C + + +
6 5 15

With the IIW equation, when C.E is less than 0.45% the steel is considered weldable with modern techniques.
The C.E of TEMPCORE reinforcing steel is well below the critical value of 0.45% and thus again is superior
to other types of reinforcing steels. The excellent weldability is also demonstrated by the tensile properties
obtained after welding. In flush butt weld no decrease in yield strength was noticed with the fracture located
outside the weld. In a BHP technical publication (1982), it was shown that under different weld and welding
processes no cracks occur in the weld. Other properties In addition to high tensile strength, excellent duc-
tility and remarkable weldability,TEMPCORE reinforcing steels exhibit good low temperature toughness, less
sensitivity to surface damage, and the fatigue resistance and sensitivity to heat are also very competitive. The
fatigue strength of TEMPCORE is as good as those of other types of reinforcing steels with equivalent yield
strength. Properties of heat resistance of TEMPCORE reinforcing steel is of importance because of the possi-
bility of fire damage. This resistance has been evaluated by two ways: tensile strength loss at room tempera-
ture after previous heat application and tensile strength loss at elevated temperature. Rehm and Russwurm
(1977) showed that after heating in laboratory conditions at temperatures between 250◦ C and 900◦ C for half
an hour, the room temperature tensile strength increases slightly with preheating up to 500◦ C and signifi-
cant drop occurs above that temperature. This property is as good as cold twisted bars and better than those
shown by some hot rolled bars. Cold -worked bars start to loss strength at 300◦ C to 400◦ C (Neves et al. 1996).
Hot rolled British III U steel loss considerable strength from 350◦ C onwards (Rehm and Russwurm 1977).
Hot-rolled low carbon micro- alloyed 400 bar starts to show loss in strength from 600◦ C. Tensile strength of
TEMPCORE reinforcing steels at elevated temperature is similar to cold worked and micro-alloyed bars with
a 20% and 40% reduction in yield strength at 300◦ C and 500◦ C respectively

2.3.6. Economic aspects


The TEMPCORE process is an economical method in producing high strength reinforcing bars. Comparing to
twisted bars, TEMPCORE process has obvious advantages in saving the cost of the mechanical twisting treat-
ment which is expensive especially for small diameter bars. A further advantage comes from reduced alloying
element requirements, off-grade heat, off-grade products, stock piling expenses and some other minor steel
making factors. The only factor which increases cost is the rolling operation related to quenching installation
and operation. Savings on alloying elements compare to non-weldable reinforcing steels amount to approxi-
mately 200 L.E/ton, but for weldable bars savings can be as high as 500 L.E/ton. When the actual composition
of the steel is considered unsuitable for its initially planned destination, it is regarded as off-grade heat. With
micro-alloyed steel, off-grade heat has to be diverted to another diameter within the same grade or another
grade. Down grade can be very costly. With TEMPCORE process, however, most of these off-grade heats can
be salvaged for the initial planned grade by adjusting the cooling power of the quenching installation. Ac-
cording to Simon et al. (1984) TEMPCORE process has brought a reduction in the percentage of the off-grade
heats by a factor of 2 to 5. TEMPCORE process provides a good quality control: with a deviation of 13.7 MPa,
12.9 MPa and 1.82% in yield strength, tensile strength and elongation, respectively TEMPCORE process has
the flexibility to produce different diameter bars and even different grades from the same steel chemistry sim-
ply by acting on the cooling power of the quenching installation. TEMPCORE process could accept scrap of
poorer quality and since the chemical composition has great flexibility, the tap-to-tap time can be reduced.
It is believed that the process is on the whole more economical than any other reinforcing bar production
method.
3
Dual Phase Steel

3.1. Introduction
3.1.1. Definition
The term dual phase steels, or DP steels, refers to a class of high strength steels which is composed of two
phases; normally a ferrite matrix and a dispersed second phase of martensite, retained austenite and/or bai-
nite. DP steels were developed in the 1970’s. The development was driven by the need for new high strength
steels without reducing the formability or increasing costs. In particular, the automotive industry has de-
manded steel grades with high tensile elongation to ensure formability, high tensile strength to establish
fatigue and crash resistance, low alloy content to ensure weldability without influencing production cost.
For years later, the demand for DP steels is still strong. Materials that can combine high strength and good
formability and thus reduce the weight of vehicles and other products give an environmental and economic
advantage. Comparing DP steels with other high strength low alloy (HSLA) steels, DP steels show superior
properties Thanks to the combination of high strength, good formability and low cost as well as high defor-

Figure 3.1: Steels for Automotive Industry

mation hardening, which implies a high energy absorbing ability or “crashworthiness”, DP steels are mainly
used by the automotive industry primarily for safety parts in car bodies, e.g. bumpers,

39
40 3. Dual Phase Steel

3.1.2. Microstructure of DP Steels


Dual-phase steels are characterized by a microstructure consisting of a fine dispersion of hard Martensite
particles in a continuous, soft, ductile Ferrite matrix. The term "dual-phase", firstly reported by Hayami &
Furukawa and thereafter adopted by the steel research community, refers to the presence of essentially two
phases, ferrite and martensite, in the microstructure, although small amounts of bainite, pearlite and retained
austenite may also be present. Irrespective of the chemical composition of the alloy, the simplest way to
obtain a dual-phase Ferritic-Martensitic steel is intercritical annealing of a Ferritic-Pearlitic microstructure
in the α+γ two-phase field, followed by a sufficiently rapid cooling to enable the austenite to Martensite
transformation (Fig. 3.2).

Figure 3.2: The microstructure of a DP800 Steel

3.1.3. DP Steel Production


Three basic approaches exist for the commercial production of dual-phase steels:
(I) As Hot – Rolled Approach
In the as-rolled process, the steel composition is chosen such that 80-90% of the steel transforms to ferrite
after the final roll pass in normal conventional hot rolling and before coiling. The remaining 10-20% does
not transform until much later, during slow cooling in the coil. This is possible with steel compositions that
exhibit certain characteristics in their continuous cooling transformation (CCT) diagram:

1. An elongated ferrite C-curve, i.e. the ability to form large amounts of ferrite.

2. A suppressed (delayed) pearlite nose and high pearlite finish temperature.

3. A gap between Pearlitic and Bainitic regions to provide temperature range.

(II) Batch or Box Annealing


The batch-annealing, where hot- or cold-rolled material is annealed in the coiled condition. Box- or batch-
annealing was mainly considered for economical and practical reasons by steel-makers where continuous
facilities were not available. Dual-phase steels could be obtained by means of batch-annealing in the in-
tercritical region for approximately 3 h to ensure homogeneity, followed by very slow cooling. Due to the
extremely low cooling rates, much higher alloying.
(III) Continuous Annealing Approach
The use of continuous annealing lines (CAL) offers the advantages of high production rates, better uniformity
of the steel properties and, furthermore, the possibility of using lower alloyed steels (having a lean chemistry).
In continuous annealing lines three types of cooling are utilized:
(a) Water – Quenching
(b) Gas – Jet Cooling
(c) Air – Cooling
The use of CAL equipped with water quenching facilities makes possible an easy and economical production
3.2. Microstructure Development during CA 41

of dual-phase high-strength steel. The major structural changes that take place in the CAL are recrystalliza-
tion and different phase transformations. During cold rolling, the ferrite grains are deformed and elongated
in the rolling direction. When heated the deformed structure starts to recrystallize and the recrystallization
start temperature is dependent on degree of deformation, chemical composition and heating rate. In the
soaking section, two main parallel processes take place; phase transformation from ferrite (α) to austenite
(γ) and carbide dissolution. The amount of austenite formed depends on the soaking temperature, the time
in the soaking section and the chemical composition of the steel. After the soaking section, the material
passes the gas-jet cooling section, where it is possible to cool the strip with gas prior to water quenching.
Even when the gas-jet cooling is turned off, the passage in the gas-jet section always involves a certain re-

transformation from austenite to new ferrite, γ α, due to temperature drop. The austenite that remains will
then transform to Martensite during water quenching. In the last section of the CAL, the reheating zone, tem-
pering of the Martensite will take place (Fig. 3.3). The following Fig. 3.5 is an example of an annealing cycle in

Figure 3.3: Schematic picture of the continuous annealing line (CAL) at SSAB in Borlänge.

CAL is presented. Indicated are, from a microstructural point of view, the major zones; preheating, soaking,
gas-jet cooling, water quenching and reheating. During the preheating process, recovery, recrystallization
and spheroidization of carbides takes place, followed by phase transformation from Ferrite (α) to Austenite
(γ) during soaking. In the gas-jet section, some formation of new ferrite occurs and the remaining austenite
transforms to Martensite during water quenching. In the reheating zone, tempering of the Martensite takes
place.

3.2. Microstructure Development during CA


3.2.1. Inheritance from the hot rolling process
The hot rolling process includes the heating of 200-220 mm thick slabs in the reheating furnaces to about
1250◦ C followed by Thermomechanical rolling in the austenitic field to a thickness of 3-5 mm. Steels for DP
production normally have a final rolling temperature of 870◦ C followed by coiling at 600◦ C. It is believed that
the coiling temperature is the process parameter in the hot rolling process having the major impact on the
mechanical properties of a cold rolled and annealed DP steel, at least when it comes to the effect of the alloy-
ing element niobium, Nb. Nb is known to be an efficient grain refining element and the most important role
of Nb as a micro alloying element during Thermomechanical rolling is the retardation of austenite recrystal-
lization. This retardation will provide more nuclei for the transformation from austenite to ferrite (γ α) and ✙
thus a finer ferrite grain size. Besides the retardation of the recrystallization processes, another important
effect of Nb is the formation of carbides and/or nitrides. The Nb(C, N) particles act as obstacles for grain
boundary migration, which leads to pancaking of the structure. The pancaking in turn provides more nuclei

for the γ α transformation and thus a smaller final ferrite grain size after hot rolling. From an equilibrium
point of view the Nb(C,N) precipitation is generally not complete; some of the Nb stays after finish rolling in
the austenite field and will effectively retard the transformation to ferrite or precipitate in the ferrite allowing
42 3. Dual Phase Steel

Figure 3.4: Production of Dual Phase Steel by Intercritical Annealing.

Figure 3.5: Production of Dual Phase Steel by Intercritical Annealing.

a strength increase by precipitation hardening.


3.3. Mechanical Properties of DP Steels 43

3.2.2. Cold Rolling


Almost all of the energy and work consumed in cold working of a metal is dissipated as heat and only a small
amount (1%) remains stored in the metal. This small amount is however the driving force for the phenomena
taking place during annealing. During cold rolling, the ferrite grains are stretched and elongated in the rolling
direction with an increase in grain boundary area as a consequence. During cold rolling, the dislocation
density increases considerably and the higher the cold rolling reduction the higher the dislocation density
and thus the driving force for recovery and recrystallization.

Figure 3.6: Microstructures of DP800 with low and high cold rolling reduction.

3.2.3. Heating & Soaking


The cold rolled material is very hard and brittle. To achieve the desired dual phase properties, the steel needs
to be annealed in the two-phase field of ferrite and austenite. Upon heating, different mechanisms control
the microstructure development dependent on chemical composition, the selected temperature and the time
needed for certain transformations (Fig. 3.7).

3.2.4. Cooling & Quenching


A reduction in temperature will affect the material in different ways depending on the cooling rate and the
chemical composition of the steel. Three microstructural changes will occur:

1. Formation of new Ferrite

2. Formation of Martensite

3. Precipitation of carbides

3.2.5. Tempering
The tempering section is the last heat treatment the strip undergoes in the continuous annealing line and the
main purpose is to reduce some of the internal stresses due to phase transformation as well as to reduce the
hardness of the Martensite.

3.3. Mechanical Properties of DP Steels


1. Forming Characteristics Dual Phase steels offer an excellent combination of strength and drawability
as a result of their good ductility and strain hardening capacity from the beginning of deformation,
which ensure homogeneous strain redistribution and reduce local thinning, no yield point elongation
(Fig. 3.8).

2. Welding Although Dual Phase steels are more highly alloyed than HSLA steels, they can be readily
welded using conventional resistance spot welding processes, provided the parameters used in indus-
trial conditions are adjusted
44 3. Dual Phase Steel

Figure 3.7: Phase transformation process in DP800, annealed in the two-phase field of α+γ.

Figure 3.8: Stress – Strain curve of Dual Phase Steel.

3. Fatigue Strength As a result of their high mechanical strength, Dual Phase steels have good fatigue
properties.

4. Impact Strength As a result of their very high tensile strength, Dual Phase steels are particularly suitable
for parts designed to absorb energy during an impact.
4
Practical Work

4.1. Practical Work Objectives


We intend to have dual phase steel composed from ferrite and martensite with higher strength, toughness
and ultimate tensile strength to yield strength ratio from ordinary rebar steel by adjusting thermal treatment
process only.

4.2. Work Procedure


The main Parameters we studied during our work in producing Dual Phase steel were as the following.

• Preheating temperature

• Rebar finishing temperature

• Quenching temperature

We study the previous parameters to control the morphology, distribution and the volume fraction of carbides
and martensite in the ferrite matrix.
General specimen preparation steps were:

• Sectioning the specimen using a saw.

• Mounting the specimen using epoxy resin.

• Grinding the specimen to remove damage introduced by sectioning.

• Grinding occurs in sequences of finer and finer abrasives.

• Typical grit sequence used 180, 320, 500, 600, 800, 1200 and finally 2000 mesh.

• Wet grinding is typically used, better to minimize heat generated.

• Polishing the specimen using alumina (Al2 O3 ) for about 40 minutes.

• Etching the specimen using 97% C2 H5 COOH and 3% (HNO3 ) solution for about 3-5 seconds and then
washing the specimen using water to remove remaining etcher then dry it by air blower.

The specimens are then ready for further steps (microscopic examination, chemical analysis and hardness
test).
the Hardness test parameters were as follows:

1. Load 500 gf

2. Dwell time 15 sec

3. Room temperature

45
46 4. Practical Work

the Equipment used to perform testing processes were the following:

• For chemical analysis we used Optical Emission Spectrometer from EZDK Company laboratories

• For the microstructure scanning we used optical microscope

• vickers hardness apparatus was used to perform the hardness tests

4.2.1. Test material


Reference specimen:
Two different grades of reinforcing steel bars with diameter 18 mm produced with TEMPCORE process at
EZDK Co. The major difference in composition between the two grades is amount of vanadium which is
very effective micro alloying element. Samples designates with the letter ’V’ have higher vanadium content
compared to those designated with the letter ’S’.

1. Sample V

• Chemical analysis

c Si Mn P S Cr Ni Cu Mo Ca
0.175 0.169 1.348 0.012 0.022 0.056 0.068 0.29 0.008742 0.0004
V Nb Ti Sn B Ceq Mn/Si% Mn/S% Ca/Al% Cr+Ni+Cu%
0.066 0.003 0.003 0.013339 0.001106 0.49429 7.966 62.559 2.049284 0.414

Table 4.1: Chemical analysis of sample V using spectrometer from EZDK labs.

• Microstructural analysis

Figure 4.1: Microstructure analysis of ’v’ sample


4.2. Work Procedure 47

Hardness profile for sample ’V’


290
Load = 500 gf
280
270
260

Hardness HV
250
240
230
220
210

−8.5 −7−6−5−4−3−2−1 0 1 2 3 4 5 6 7 8.5

• Hardness profile Distance from the core (mm)

2. Sample S

• Chemical analysis

c Si Mn P S Cr Ni Cu Mo Ca
0.389 0.188 0.931 0.013 0.032 0.074 0.067 0.273 0.010169 0.0009
V Nb Ti Sn B Ceq Mn/Si% Mn/S% Ca/Al% Cr+Ni+Cu%
0.004 0.002 0.003 0.014147 0.001427 0.615789 4.959 28.648 4.488508 0.414

Table 4.2: Chemical analysis of sample V using spectrometer from EZDK labs.

• Microstructure analysis

Figure 4.2: Microstructure analysis of ’s’ sample


48 4. Practical Work

Hardness profile for sample ’V’


290
Load = 500 gf
280
270
260

Hardness HV
250
240
230
220
210

−8.5 −7−6−5−4−3−2−1 0 1 2 3 4 5 6 7 8.5

• Hardness profile Distance from the core (mm)

Fig. 4.2 & Fig. 4.1 show the variation of hardness across the three layers. The hardness is lowest in the central
core and increases gradually toward the edges as expected. The soft inner core is composed of ferrite and
pearlite, whereas the outer surface hard layer is made up of tempered martensite and these two regions are
separated by the transition zone which is a thin layer of bainite or mixture of bainite and pearlite depending
on cooling rate.
From chemical analysis we were able to calculate critical transformation temperatures (e.g. Ae 3 , Ae 1 , M s and
B s ) using the following empirical equations.

Empirical equations of calculating critical temperatures:

M sT M = 420 − 208.33C − 33.428Mn + 1.296Mn 2 − 0.02167Mn 3 − 16.08N i + 0.7817N i 2


− 0.0246N i 3 − 2.473C r + 30.00Mo + 12.86C o − 0.2654C o 2 + 0.001547C o 3 (4.1)
2 2 3
− 7.18Cu − 72.65N − 43.36N − 16.28Ru − 1.72Ru − 0.08117Ru

M sLM = 540 − 356.25C − 47.59Mn + 2.25Mn 2 − 0.0415Mn 3 − 24.65N i + 1.36N i 2 − 0.0384N i 3 − 17.82C r
+ 1.42C r 2 + 17.50Mo + 21.87C o − 0.468C o 2 + 0.00296C o 3 − 16.52Cu − 260.64N − 17.66Ru
(4.2)

B s = 630 − 45Mn − 40V − 35Si − 30C r − 25Mo − 20N i − 15W (4.3)

A e1 = 723 − 16.9N i + 29.1Si + 6.38W − 10.7Mn + 16.9C r + 290As (4.4)

p
A e3 = 910 − 203 C + 44.7S − 15.2N + 31.5Mo + 104V + 13.1W (4.5)
− 30Mn + 11C r + 20Cu − 700P − 400Al − 120As − 400T i

Notation:
Ae 1 : Lower Equilibrium Temperature between Ferrite and Austenite [◦ C]
Ae 3 : Upper Equilibrium Temperature between Ferrite and Austenite [◦ C]
B s : Bainite Start Temperature [◦ C]
M sT M : Twinned Martensite Start Temperature [◦ C]
M sLM : Lath Martensite Start Temperature [◦ C]
Alloy Amount: [weight %]
4.2. Work Procedure 49

The critical temperatures were found to be:

Critical temp [◦ C] Ae 1 Ae 3 Bs Ms
Sample V 713 827.5 503.1 410.2
Sample S 718.5 774.6 517.3 360

Table 4.3: Critical temperatures calculated using empirical formulae.

To ensure that the prior deformation of specimen will not interfere the heat-treatment process we re-
moved deformation history of specimens using ordinary normalizing treatment
Normalizing steps:

1. Heating the specimens above their respective A 3 temperatures for 5 minutes.

2. Specimens then were allowed to cool down to room temperature in still air.

3. we planned to repeat testing procedure to assure that the history was successfully removed but unfor-
tunately we couldn’t do that due to the quarantine imposed by the government.

Some specimens were heat treated randomly to define the upper and the lower boundaries of mechanical
properties that we can achieve using heat-treatment of these grades of steel, the treatments are listed in (Ta-
ble 4.4).

Specimen code Quenching temperature(◦ C) Holding time at quenching temperature (min)


So 740 5
Vo 740 5
SI 740 15
VI 740 15
SR 760 5
VR 760 5
SH 760 15
VH 760 15

Table 4.4: Heat Treatment Operations Carried out.

based on the Hardness profiles shown above above - hardness profiles of sample ’V’ and sample ’S’ - and
on microstructures of the two specimens shown in Fig. 4.2 & Fig. 4.1 it is only possible to observe the effect
of steel chemical composition difference on both the hardness distribution over the cross section of treated
specimens and the microstructures. From the hardness distribution curves an almost similar values are ob-
tained for both composition steels with a higher surface hardness at the outer surface for the sample with
V-addition. Moreover, a more refined microstructure of both temper martensite zone and ferrite and pearlite
core zone, such refinement could also attributed to chemical composition mainly the V- addition compared
with V- free sample. the specimens in the previous table and the heat treatments they went through was
meant to be a guide helping us explore the variations of mechanical properties of the specimens with pro-
cessing parameters thus allowing us to narrow down our search for the optimum results also we intended to
draw out the resemblance and matches between our work and theoretical models based on thermodynamic
modeling and simulations of phase transformations, which in turn allows us to better understand how the
dual phase microstructure develops and develop processing-structure-property linkages. Unfortunately we
couldn’t complete the laboratory work due to corona virus pandemic and quarantine.
Bibliography
[1] Harshad Bhadeshia and Robert Honeycombe. Iron-carbon equilibrium and plain carbon steels. In Steels:
Microstructure and Properties, pages 59–100. Elsevier, 2017.

[2] F.C. Campbell, editor. Elements of Metallurgy and Engineering Alloys. ASM International, 2008.

[3] Elena Pereloma and David V. Edmonds. Phase transformations in steels. Woodhead Publishing Limited,
2012.

[4] Hany Khalifa, G. M. Megahed, Rawia M. Hamouda, and Mohamed A. Taha. Experimental investigation
and simulation of structure and tensile properties of tempcore treated rebar. Journal of Materials Pro-
cessing Technology, 230:244–253, apr 2016.

[5] O. Niño, D. Martínez, C. Lizcano, Martha Patrizia Guerrero-Mata, and Rafael Colás. Study of the tem-
pcore process for the production of high resistance reinforcing rods. Materials Science Forum, 537-
538:533–540, feb 2007.

[6] Pierre Simon, Mario Economopoulos, Paul Nilles, et al. Tempcore, an economical process for the pro-
duction of high quality rebars. Metallurgical Plant and Technology, 96(3):80–93, 1984.

[7] G Rehm and D Russwurm. Assessment of concrete reinforcing bars made by the tempcore process. C.
R. M. Metall. Rep., 31(51):3–16, 1977.

[8] B. Hortigón, F. Ancio, E. J. Nieto-García, M. A. Herrera, and J. M. Gallardo. Influence of rebar design on
mechanical behaviour of tempcore steel. Procedia Structural Integrity, 13:601–606, 2018.

[9] Hakan Çetinel. An investigation of the effect of variable parameters on the material quality in tempcore
process. PhD thesis, DEÜ Fen Bilimleri Enstitüsü, 1997.

[10] Barbara Maffei, Walter Salvatore, and Renzo Valentini. Dual-phase steel rebars for high-ductile r.c. struc-
tures, part 1: Microstructural and mechanical characterization of steel rebars. Engineering Structures,
29(12):3325–3332, dec 2007.

[11] Zhonghao Jiang, Zhenzhong Guan, and Jianshe Lian. Effects of microstructural variables on the defor-
mation behaviour of dual-phase steel. Materials Science and Engineering: A, 190(1-2):55–64, 1995.

[12] Silvia Caprili, Walter Salvatore, Renzo Valentini, Cristiano Ascanio, and Gianbruno Luvarà. Dual-phase
steel reinforcing bars in uncorroded and corroded conditions. Construction and Building Materials,
218:162–175, 2019.

[13] D. T. Llewellyn and R. C. Hudd. Low-carbon structural steels. In Steels, pages 137–198. Elsevier, 1998.

[14] Ylva Granbom. Effects of process parameters prior to annealing on the formability of two cold rolled
dual phase steels. steel research international, 79(4):297–305, apr 2008.

[15] Nina Fonstein. Effect of structure on mechanical properties of dual-phase steels. In Advanced High
Strength Sheet Steels, pages 67–138. Springer International Publishing, 2015.

[16] Y. Granbom. Influence of niobium and coiling temperature on the mechanical properties of a cold rolled
dual phase steel. Revue de Métallurgie, 104(4):191–197, apr 2007.

[17] P. E. Repas. Metallurgy, production technology, and properties of dual-phase sheet steels. In SAE Tech-
nical Paper Series. SAE International, feb 1979.

[18] Nina Fonstein. The effect of chemical composition on formation of ferrite–martensite structures and
mechanical properties of dual-phase steels. In Advanced High Strength Sheet Steels, pages 139–184.
Springer International Publishing, 2015.

51

You might also like